| 0.0.0.0 | Abstract |
| 0.0.0.0 | Abstract | | | | The global methane (CH4) budget is becoming an increasingly
important component for managing realistic pathways to mitigate climate change.
This relevance, due to a shorter atmospheric lifetime and a stronger warming potential
than carbon dioxide, is challenged by the still unexplained changes of
atmospheric CH4 over the past decade. Emissions and concentrations
of CH4 are continuing to increase, making CH4 the second
most important human-induced greenhouse gas after carbon dioxide. Two major
difficulties in reducing uncertainties come from the large variety of diffusive
CH4 sources that overlap geographically, and from the destruction of
CH4 by the very short-lived hydroxyl radical (OH). To address these
difficulties, we have established a consortium of multi-disciplinary scientists
under the umbrella of the Global Carbon Project to synthesize and stimulate
research on the methane cycle, and producing regular (∼
biennial) updates of the global methane budget. This consortium includes
atmospheric physicists and chemists, biogeochemists of surface and marine
emissions, and socio-economists who study anthropogenic emissions. Following
Kirschke et al. (2013), we propose here the first version of a living review
paper that integrates results of top-down studies (exploiting atmospheric
observations within an atmospheric inverse-modelling framework) and bottom-up
models, inventories and data-driven approaches (including process-based models
for estimating land surface emissions and atmospheric chemistry, and
inventories for anthropogenic emissions, data-driven extrapolations).
For the 2003–2012 decade, global methane emissions are estimated by top-down
inversions at 558 Tg CH4 yr-1, range 540–568. About 60 %
of global emissions are anthropogenic (range 50–65 %). Since 2010, the
bottom-up global emission inventories have been closer to methane emissions in
the most carbon-intensive Representative Concentrations Pathway (RCP8.5) and
higher than all other RCP scenarios. Bottom-up approaches suggest larger global
emissions (736 Tg CH4 yr-1, range 596–884) mostly because
of larger natural emissions from individual sources such as inland waters,
natural wetlands and geological sources. Considering the atmospheric
constraints on the top-down budget, it is likely that some of the individual
emissions reported by the bottom-up approaches are overestimated, leading to
too large global emissions. Latitudinal data from top-down emissions indicate a
predominance of tropical emissions (∼ 64 % of the global budget,
< 30◦ N) as compared to mid (∼ 32 %, 30–60◦ N) and
high northern latitudes (∼ 4 %, 60–90◦ N).
Top-down inversions consistently infer lower emissions in China (∼
58 Tg CH4 yr-1, range 51–72, −14 %) and higher
emissions in Africa (86 Tg CH4 yr-1, range 73–108,
>+19 %) than bottom-up values used as prior estimates. Overall,
uncertainties for anthropogenic emissions appear smaller than those from
natural sources, and the uncertainties on source categories appear larger for
top-down inversions than for bottom-up inventories and models.
The most important source of uncertainty on the methane budget is
attributable to emissions from wetland and other inland waters. We show that
the wetland extent could contribute 30–40 % on the estimated range for wetland
emissions. Other priorities for improving the methane budget include the
following: (i) the development of process-based models for inland-water
emissions, (ii) the intensification of methane observations at local scale
(flux measurements) to constrain bottom-up land surface models, and at regional
scale (surface networks and satellites) to constrain top-down inversions, (iii)
improvements in the estimation of atmospheric loss by OH, and (iv) improvements
of the transport models integrated in top-down inversions. The data presented
here can be downloaded from the Carbon Dioxide Information Analysis Center (http://doi.org/10.3334/CDIAC/GLOBAL_
METHANE_BUDGET_2016_V1.1) and the Global Carbon Project.
|
|
| 0.0.0.0 | Copyright statement |
| 0.0.0.0 | Copyright statement | | | | The works published in this journal are distributed under the Creative
Commons Attribution 3.0 License. This license does not affect the Crown
copyright work, which is re-usable under the Open Government Licence (OGL). The
Creative Commons Attribution 3.0 License and the OGL are interoperable and do
not conflict with, reduce or limit each other.
|
|
| 1.0.0.0 | Introduction |
| 1.0.0.0 | Introduction | | | | The surface dry air mole fraction of atmospheric methane (CH4)
reached 1810 ppb in 2012 (Fig. 1). This level, 2.5 times larger than in 1750,
results from human activities related to agriculture (livestock, rice
cultivation), fossil fuel usage and waste sectors, and from climate and CO2-changes
affecting natural emissions (Ciais et al., 2013). Atmospheric CH4 is
the second most impactful anthropogenic greenhouse gas after carbon dioxide (CO2)
in terms of radiative forcing. Although its global emissions, estimated at
around 550 Tg CH4 yr-1 (Kirschke et al., 2013), are only
4 % of the global CO2 anthropogenic emissions in units of carbon
mass flux, atmospheric CH4 has contributed 20 % (∼
0.48 W m-2,) of the additional radiative forcing accumulated in the
lower atmosphere since 1750 (Ciais et al., 2013).
This is because of the larger warming potential of methane compared to CO2,
about 28 times on a 100-year horizon as re-evaluated by the Intergovernmental
Panel on Climate Change (IPCC) 5th Assessment Report (AR5) (when using Global
Warming Potential metric; Myhre et al., 2013). Changes in other chemical
compounds (such as NOxor CO) also influence the forcing of
methane through changes in its lifetime. From an emission point of view, the
radiative impact attributed to CH4 emissions is about 0.97 W m.
This is because emission of CH4 leads to production of ozone, of
stratospheric water vapour, and of CO2, and importantly affects its
own lifetime (Myhre et al., 2013; Shindell et al., 2012). CH4 has a
short lifetime in the atmosphere (∼9 years for the modern
inventory; Prather et al., 2012), and a stabilization or reduction of CH4
emissions leads rapidly to a stabilization or reduction of methane radiative
forcing. Reduction in CH4 emissions is therefore an effective option
for climate change mitigation. Moreover, CH4 is both a greenhouse
gas and an air pollutant, and as such covered by two international conventions:
the United Nations Framework Convention on Climate Change (UNFCCC) and the
Convention on Long Range Transport of Air Pollution (CTRTAP).
Changes in the magnitude and timing (annual to interannual) of individual
methane sources and sinks over the past decades are uncertain (Kirschke et al.,
2013) with relative uncertainties (hereafter reported as min–max ranges) of 20–
30 % for inventories of anthropogenic emissions in each sector (agriculture,
waste, fossil fuels) and for biomass burning, 50 % for natural wetland
emissions and reaching 100 % or more for other natural sources (e.g. inland
waters, geological). The uncertainty in the global methane chemical loss by OH,
the predominant sink, is estimated between 10 % (Prather et al., 2012) and 20 %
(Kirschke et al., 2013), implying a similar uncertainty in global methane
emissions as other sinks are much smaller and the atmospheric growth rate is
well defined (Dlugokencky et al., 2009). Globally, the contribution of natural
emissions to the total emissions is reasonably well quantified by combining
lifetime estimates with reconstructed preindustrial atmospheric methane
concentrations from ice cores (e.g. Ehhalt et al., 2001). Uncertainties in
emissions reach 40–60 % at regional scale (e.g. for South America, Africa,
China and India). Beyond the intrinsic value of characterizing the
biogeochemical cycle of methane, understanding the evolution of the methane
budget has strong implications for future climate emission scenarios. Worryingly,
the current emission trajectory is tracking the warmest of all IPCC scenarios,
the RCP8.5, and is clearly inconsistent with lower temperature scenarios, which
show substantial to large reductions of methane emissions (Collins et al.,
2013).
Reducing uncertainties in individual methane sources, and thus in the
overall methane budget, is not an easy task for, at least, four reasons. First,
methane is emitted by a variety of processes that need to be understood and
quantified separately, both natural or anthropogenic, point or diffuse sources,
and associated with three main emission processes (biogenic, thermogenic and
pyrogenic). Among them, several important anthropogenic CH4 emission
sources are poorly reported. These multiple sources and processes require the
integration of data from diverse scientific communities to assess the global
budget. Second, atmospheric methane is removed by chemical reactions in the
atmosphere involving radicals (mainly OH), which have very short lifetimes
(typically 1 s). Although OH can be measured locally, its spatiotemporal
distribution remains uncertain at regional to global scales, which cannot be
assessed by direct measurements. Third, only the net methane budget (sources –
sinks) is constrained by the precise observations of the atmospheric growth
rate (Dlugokencky et al., 2009), leaving the sum of sources and the sum of
sinks uncertain. One simplification for CH4 compared to CO2
is that the oceanic contribution to the global methane budget is very small (∼
1–3 %), making source estimation mostly a continental problem (USEPA, 2010a).
Finally, we lack observations to constrain (1) process models that produce
estimates of wetland extent (Stocker et al., 2014; Kleinen et al., 2012) and
emissions (Melton et al., 2013; Wania et al., 2013), (2) other inland water
sources (Bastviken et al., 2011), (3) inventories of anthropogenic emissions
(USEPA, 2012; EDGARv4.2FT2010, 2013), and (4) atmospheric inversions, which aim
at representing or estimating the different methane emissions from global to
regional scales (Houweling et al., 2014; Kirschke et al., 2013; Bohn et al.,
2015; Spahni et al., 2011; Tian et al., 2016). Finally, information contained
in the ice core methane records has only been used in a few studies to evaluate
process models (Zürcher et al., 2013; Singarayer et al., 2011).
The regional constraints brought by atmospheric sampling on atmospheric
inversions are significant for northern mid latitudes thanks to a number of
high-precision and highaccuracy surface stations (Dlugokencky et al., 2011).
The atmospheric observation density has improved in the tropics with
satellite-based column-averaged methane mixing ratios (Buchwitz et al., 2005b;
Frankenberg et al., 2005; Butz et al., 2011). However, the optimal usage of satellite
data remains limited by systematic errors in satellite retrievals (Bergamaschi
et al., 2009; Locatelli et al., 2015). The development of low-bias observations
system from space, such as active lidar technics, is promising to overcome
these issues (Kiemle et al., 2014). The partition of regional emissions by
processes remains very uncertain today, waiting for the development or
consolidation of measurements of more specific tracers, such as methane
isotopes or ethane, dedicated to constrain the different methane sources or
groups of sources (e.g. Simpson et al., 2012; Schaefer et al., 2016; Hausmann
et al., 2016).
The Global Carbon Project (GCP) aims at developing a complete picture of the
carbon cycle by establishing a common, consistent scientific knowledge to
support policy debate and actions to mitigate the rate of increase of
greenhouse gases in the atmosphere (http://www.globalcarbonproject. org). The
objective of this paper is to provide an analysis and synthesis of the current
knowledge about the global and regional methane budgets by gathering results of
observations and models and by extracting from these the robust features and
the uncertainties remaining to be addressed. We combine results from a large
ensemble of bottom-up approaches (process-based models for natural wetlands,
data-driven approaches for other natural sources, inventories of anthropogenic
emissions and biomass burning, and atmospheric chemistry models) and of
top-down approaches (methane atmospheric observing networks, atmospheric
inversions inferring emissions and sinks from atmospheric observations and
models of atmospheric transport and chemistry). The focus here is on decadal
budgets, leaving in-depth analysis of trends and year-to-year changes to future
publications. This paper is built on the principle of a living review to be
published at regular intervals (e.g. every two years) and will synthesize and
update new annual data, the introduction of new data products, model
development improvements, and new modelling approaches to estimate individual
components contributing to the CH4 budget.
The work of Kirschke et al. (2013) was the first GCP-like CH4
budget synthesis. Kirschke et al. (2013) reported decadal mean CH4
emissions and sinks from 1980 to 2009 based on bottom-up and top-down
approaches. Our new analysis, and our approach for the living review budget,
will report methane emissions for three targeted time periods:
the last calendar decade (2000–2009, for this paper),
the last available decade (2003–2012, for this paper), and
the last available year (2012, for this paper). Future efforts will also
focus on retrieving budget data as recent as possible.
Five sections follow this introduction. Section 2 presents the methodology
to treat and analyse the data streams. Section 3 presents the current knowledge
about methane sources and sinks based on the ensemble of bottom-up approaches
reported here (models, inventories, data-driven approaches). Section 4 reports
the atmospheric observations and the top-down inversions gathered for this
paper. Section 5, based on Sects. 3 and 4, provides an analysis of the global
methane budget (Sect. 5.1) and of the regional methane budget (Sect. 5.2).
Finally Sect. 6 discusses future developments, missing components and the
largest remaining uncertainties after this update on the global methane budget.
|
|
| 2.0.0.0 | Methodology |
| 2.0.0.0 | Methodology | | | | Unless specified, the methane budget is presented in teragrammes of CH4
per year (1 Tg CH4 yr-1 >= 1012 g CH4 yr-1),
methane concentrations as dry air mole fractions in parts per billion (ppb) and
the methane annual increase, GATM, in ppb yr-1. In the
different tables, we present mean values and ranges for the last calendar
decade (2000–2009, for this paper), the period 2003–2012, together with results
for the last available year (2012, for this paper). Results obtained from the
previous synthesis are also given (Kirschke et al., 2013, for this paper).
Following Kirschke et al. (2013) and considering the relatively small and
variable number of studies generally available for individual numbers,
uncertainties are reported as minimum and maximum values of the gathered
studies in brackets. In doing so, we acknowledge that we do not take into
account all the uncertainty of the individual estimates (when provided). This
means that the full uncertainty range may be greater than the range provided
here. These minimum and maximum values are those calculated using the boxplot
analysis presented below and thus excluding identified outliers when existing.
The CH4 emission estimates reported in this paper, derived mainly
from statistical calculations, are given with up to three digits for
consistency across all budget flux components and to ensure conservation of
quantities when aggregated into flux categories in Table 2 (and regional
sources in Table 4). However, the reader should keep in mind the associated
uncertainties and acknowledge a two-digit global methane budget.
|
|
| 2.1.0.0 | Processing of emission maps |
| 2.1.0.0 | Processing of emission maps | | | | Common data analysis procedures have been applied to the different bottom-up
models, inventories and atmospheric inversions whenever gridded products exist.
The monthly or yearly fluxes (emissions and sinks) provided by different groups
were processed similarly. They were re-gridded on a common grid (1◦ ×
1◦) and converted into the same units (Tg CH4
per grid cell). For coastal pixels of land fluxes, to avoid allocating land
emissions into oceanic areas when regridding the model output, all emissions
were re-allocated to the neighbouring land pixel. The opposite was done for
ocean fluxes. Monthly, annual and decadal means were computed from the gridded
1◦ by 1◦ maps.
|
|
| 2.2.0.0 | Definition of the boxplots |
| 2.2.0.0 | Definition of the boxplots | | | | Most budgets are presented as boxplots, which have been created using
routines in IDL language, provided with the standard version of the IDL
software. The values presented in the following are calculated using the
classical conventions of boxplots including quartiles (25 %, median, 75 %),
outliers, and minimum and maximum values (without the outliers). Outliers are
determined as values below the first quartile minus 3 times the interquartile
range or values above third quartile plus 3 times the interquartile range.
Identified outliers (when existing) are plotted as stars on the different
figures proposed. The mean values are reported in the tablesand represented as
“>+” symbols in the figures.
|
|
| 2.3.0.0 | Definition of regions and source categories |
| 2.3.0.0 | Definition of regions and source categories | | | | Geographically, emissions are reported for the global scale, for three
latitudinal bands (< 30, 30–60, 60–90◦ N, only for gridded products)
and for 15 regions (oceans and 14 continental regions, see Sect. 5 and Fig. 7
for region map). As anthropogenic emissions are reported at country level, we
chose to define the regions based on a country list (Supplement Table S1). This
approach is compatible with all topdown and bottom-up approaches providing
gridded products as well. The number of regions was chosen to be close to the
widely used TransCom intercomparison map (Gurney et al., 2004), but with
subdivisions to isolate important countries for the methane budget (China,
India, USA and Russia). Therefore, the new region map defined here is different
from the TransCom map but more adapted to the methane cycle. One caveat is that
the regional totals are not directly comparable with other studies reporting
methane emissions on the TransCom map (as in Kirschke et al., 2013, for
example), although the names of some regions are the same.
Bottom-up estimates of methane emissions rely on models for individual
processes (e.g. wetlands) or on inventories representing different source types
(e.g. gas emissions). Chemistry transport models generally represent methane
sinks individually in their chemical schemes (Williams et al., 2012).
Therefore, it is possible to represent the bottom-up global methane budget for
all individual sources. However, by construction, the total methane emissions
derived from a combination of independent bottom-up estimates are not
constrained.
For atmospheric inversions (top-down), the situation is different.
Atmospheric observations provide a constraint on the global source, given a
fairly strong constraint on the global sink derived using a proxy tracer such
as methyl chloroform (Montzka et al., 2011). The inversions reported in this
work solve either for a total methane flux (e.g. Pison et al., 2013) or for a
limited number of flux categories (e.g. Bergamaschi et al., 2013). Indeed, the
assimilation of CH4 observations alone, as reported in this
synthesis, cannot fully separate individual sources, although sources with
different locations or temporal variations could be resolved by the assimilated
atmospheric observations. Therefore, following Kirschke et al. (2013), we have
defined five broad categories for which top-down estimates of emissions are
given: natural wetlands, agriculture and waste emissions, fossil fuel
emissions, biomass and biofuel burning emissions, and other natural emissions
(other inland waters, wild animals, wildfires, termites, land geological
sources, oceanic sources (geological and biogenic), and terrestrial
permafrost). Global and regional methane emissions per source category were
obtained directly from the gridded optimized fluxes if an inversion solved for
the GCP categories. Alternatively, if the inversion solved for total emissions
(or for different categories embedding GCP categories), then the prior
contribution of each source category at the spatial resolution of the inversion
was scaled by the ratio of the total (or embedding category) optimized flux
divided by the total (or embedding category) prior flux (Kirschke et al.,
2013). Also, the soil uptake was provided separately in order to report the
total surface emissions and not net emissions (sources minus soil uptake). For
bottom-up, some individual sources can be found gridded in the literature
(anthropogenic emissions, natural wetlands), but some others are not gridded
yet (e.g. inland waters, geological, oceanic sources). The regional bottom-up
methane budget per source category is therefore presented only for gridded categories
(all but the “other natural” category).
In summary, bottom-up models and inventories are presented for all
individual sources and for the five broad categories defined above at global
scale, and only for four broad categories at regional scale. Top-down
inversions are reported globally and regionally for the five broad categories
of emissions.
|
|
| 3.0.0.0 | Methane sources and sinks |
| 3.0.0.0 | Methane sources and sinks | | | | Here we provide a complete review of all methane sources and sinks based on
an ensemble of bottom-up approaches from multiple sources: process-based
models, inventories, and data-driven methods. For each source, a description of
the involved emitting process(es) is given, together with a brief description
of the original datasets (measurements, models) and the related methodology. Then,
the estimate for the global source and its range is given and analysed.
Detailed descriptions of the datasets can be found elsewhere (see references of
each component in the different subsections and tables).
Methane is emitted by a variety of sources in the atmosphere. These can be
sorted by emitting process (thermogenic, biogenic or pyrogenic) or by
anthropogenic vs. natural origin. Biogenic methane is the final product of the
decomposition of organic matter by Archaea
in anaerobic environments, such as water-saturated soils, swamps, rice
paddies, marine sediments, landfills, waste-water facilities, or inside animal
intestines. Thermogenic methane is formed on geological timescales by the
breakdown of buried organic matter due to heat and pressure deep in the Earth’s
crust. Thermogenic methane reaches the atmosphere through marine and land
geologic gas seeps and during the exploitation and distribution of fossil fuels
(coal mining, natural gas production, gas transmission and distribution, oil
production and refinery). Finally, pyrogenic methane is produced by the
incomplete combustion of biomass. Peat fires, biomass burning in deforested or
degraded areas, and biofuel usage are the largest sources of pyrogenic methane.
Methane hydrates, icelike cages of methane trapped in continental shelves and
below sub-sea and land permafrost, can be of biogenic or thermogenic origin.
Each of the three process categories has both anthropogenic and natural
components. In the following, we choose to present the different methane
sources depending on their anthropogenic or natural origin, which seems more
relevant for planning climate mitigation activities. However this choice does
not correspond exactly to the definition of anthropogenic and natural used by
UNFCCC and IPCC guidelines, where, for pragmatic reasons, all emissions from
managed land are reported as anthropogenic, which is not the case here. For
instance, we consider all wetlands in the natural emissions whereas there are
managed wetlands.
|
|
| 3.1.0.0 | Anthropogenic methane sources |
| 3.1.0.0 | Anthropogenic methane sources | | | | Various human activities lead to the emissions of methane to the atmosphere.
Agricultural processes under anaerobic conditions such as wetland rice
cultivation and livestock (enteric fermentation in animals, and the
decomposition of animal wastes) emit biogenic CH4, as does the
decomposition of municipal solid wastes. Methane is also emitted during the
production and distribution of natural gas and petroleum and is released as a
byproduct of coal mining and incomplete fossil fuel and biomass combustion
(USEPA, 2016).
Emission inventories were developed to generate bottom-up estimates of
sector-specific emissions by compiling data on human activity levels and
combining them with the associated emission factors.
An ensemble of individual inventories was gathered here to estimate
anthropogenic methane emissions. We also refer to the extensive assessment
report of the Arctic Monitoring and Assessment Programme (AMAP) published in
2015 on “Methane as Arctic climate forcer” (Höglund-Isaksson et al., 2015),
which provides a detailed presentation and description of methane inventories
and global scale estimates for the year 2005 (see their chap. 5 and in
particular their Tables 5.1 to 5.5).
|
|
| 3.1.1.0 | Reported global inventories |
| 3.1.1.0 | Reported global inventories | | | | The main three bottom-up global inventories covering all anthropogenic
emissions are from the United States Environmental Protection Agency, USEPA
(2012, 2006), the Greenhouse gas and Air pollutant Interactions and Synergies
(GAINS) model developed by the International Institute for Applied Systems
Analysis (IIASA) (Höglund-Isaksson, 2012) and the Emissions Database for Global
Atmospheric Research (EDGARv4.1, 2010; EDGARv4.2FT2010, 2013).
The latter is an inventory compiled by the European Commission Joint
Research Centre (EC-JRC) and Netherland’s Environmental Assessment Agency
(PBL). These inventories report the major sources of anthropogenic methane
emissions: fossil fuel production, transmission and distribution; livestock
(enteric fermentation and manure management); rice cultivation; solid waste and
waste water. However, the level of detail provided by country and by sector
varies between inventories, as these inventories do not consider the same
number of geographical regions and source sectors (Höglund-Isaksson et al.,
2015, see their Table 5.2). In these inventories, methane emissions for a given
region/country and a given sector are usually calculated as the product of an
activity level, an emission factor for this activity and an abatement
coefficient to account for regulations implemented to control emissions if
existing (see Eq. 5.1 of Höglund-Isaksson et al., 2015; IPCC, 2006). The
integrated emission models USEPA and GAINS provide estimates every 5 or 10
years for both past and future periods, while EDGAR provides annual estimates
only for past emissions. These datasets differ in their assumptions and the
data used for the calculation; however, they are not completely independent as
they follow the IPCC guidelines (IPCC, 2006). While the USEPA inventory adopts
the emissions reported by the countries to the UNFCCC, EDGAR and the GAINS
model produced their own estimates using a consistent approach for all
countries. As a result, the latter two approaches need large country-specific
information or, if not available, they adopt IPCC default factors or emission
factors reported to UNFCCC (Olivier et al., 2012; Höglund-Isaksson, 2012).
Here, we also integrate the Food and Agriculture Organization (FAO) dataset,
which provides estimates of methane emissions at country level but only for
agriculture (enteric fermentation, manure management, rice cultivation, energy
usage, burning of crop residues and of savannahs) and land use (biomass
burning) (FAO, 2016). It will hereafter be referred as FAO-CH4.
FAO-CH4 uses activity data from the FAOSTAT database as reported by
countries to National Agriculture Statistical Offices (Tubiello et al., 2013)
and mostly the Tier 1 IPCC methodology for emission factors (IPCC, 2006), which
depend on geographic location and development status of the country. For
manure, the necessary country-scale temperature was obtained from the FAO
global agro-ecological zone database (GAEZv3.0, 2012).
We use the following versions of these inventories: version EDGARv4.2FT2010
that provides yearly gridded emis sions by sectors from 2000 to 2010 (Olivier
and Janssens-Maenhout, 2012; EDGARv4.2FT2010, 2013), version 5a of the GAINS
model (Höglund-Isaksson, 2012) that assumes current legislation for air
pollution for the future, the revised estimates of 2012 from the USEPA (2012),
and finally the FAO emission database accessed in April 2016. Further details
of the inventories used in this study are provided in Table 1. Overall, only
EDGARv4.2FT2010 and GAINS provide gridded emission maps by sectors, and only
EDGAR provides gridded maps on a yearly basis, which ex plains why this
inventory is the most used in inverse mod elling. These inventories are not all
regularly updated. For the purpose of this study, the estimates from USEPA and
GAINS have been linearly interpolated to provide yearly values, as provided by
the EDGAR inventory. We also use the EDGARv4.2FT2012 data, which is an update
of the time series of the country total emissions until 2012 (Rogelj et al.,
2014; EDGARv4.2FT2012, 2014). This update has been developed based on
EDGARv4.2FT2010 and uses IEA energy balance statistics (IEA, 2013) and NIR/CRF
of UNFCCC (2013), as described in part III of IEA’s CO2 book by Olivier and
Janssens-Maenhout (2014).
For this study, engaged before the update of EDGARv4.2 up to 2012, we built
our own update from 2008 up to 2012 using FAO emissions to quantify CH4
emissions from enteric fermentation, manure management and rice cultivation
(described above) and BP statistical review of fossil fuel production and
consumption (http://www.bp.com/)
to update CH4 emissions from coal, oil and gas sectors. In this
inventory, called EDGARv4.2EXT, methane emissions after 2008 are set up equal
to the FAO emissions (or BP statistics) of year t times the ratio between the
mean EDGAR CH4 emissions (EEDGARv4.2)
over 2006–2008 and the mean value of FAO emissions (VFAOin the
following equation) (or BP statistics) over 2006–2008. For each emission
sector, the country-specific emissions (EEDGARv4.2ext)
in year (t )
are estimated following Eq. (1):(See original document for equation)
Other sources than those aforementioned are kept constant at the 2008 level.
This extrapolation approach is necessary and often performed by top-down
inversions to define prior emissions, because, up to now, global inventories
such as sector-specific emissions in EDGAR database have not been updated on a
regular basis. EC-JRC released, however, their update up to 2012
(EDGARv4.2FT2012) containing country total emissions, which allows evaluation
of our extrapolation approach. The extrapolated global totals of EDGARv4.2EXT
are within 1 % of EDGARv4.2FT2012.
3.1.2 |
|
| 3.1.2.0 | Total anthropogenic methane emissions |
| 3.1.2.0 | Total anthropogenic methane emissions | | | | Based on the ensemble of inventories detailed above, anthr3.1.3opogenic
emissions are ∼ 352 [340–360] Tg CH4 yr-1
for the decade 2003–2012 (Table 2, including biomass and biofuel burning). For
the 2000–2009 period, anthropogenic emissions are estimated at ∼
338 [329–342] Tg CH4 yr-1. This estimate is consistent,
albeit larger and with a smaller uncertainty range than Kirschke et al. (2013)
for the 2000–2009 decade (331 Tg CH4 yr-1 [304–368]).
Such differences are due to the different sets of inventories gathered. The
range of our estimate (∼ 5 %) is smaller then the
range reported in the AMAP assessment report (∼ 20 %) both because the
latter was reporting more versions of the different inventories and
projections, and because it was for the particular year 2005 and not for a
decade as here.
Figure 2 presents the global methane emissions of anthropogenic sources
(excluding biomass and biofuel burning) estimated and projected by the
different inventories between 2000 and 2020. The inventories consistently
estimate that about 300 Tg of methane was released into the atmosphere in 2000 by
anthropogenic activities. The main discrepancy between the inventories is
observed in their trend after 2005 with the lowest emissions projected by USEPA
and the largest emissions estimated by EDGARv4.2FT2012. The increase in CH4
emissions is mainly determined from coal mining, whose activity increased
considerably in China from 2002 to 2012 (see Sect. 3.1.3).
Despite relatively good agreement between the inventories on total emissions
from year 2000 onwards, large differences can be found at the sector and
country levels (IPCC, 2014). Some of these discrepancies are detailed in the
following sections.
For the fifth IPCC Assessment Report, four representative concentration
pathways (RCPs) were defined RCP8.5, RCP6, RCP4.5 and RCP2.6 (the latter is
also referred to as RCP3PD, where “PD” stands for peak and decline). The
numbers refer to the radiative forcing by the year 2100 in W m-2,.
These four independent pathways developed by four individual modelling groups
start from the identical base year 2000 (Lamarque et al., 2010) and have been
harmonized with historical emissions up to 2005. An interesting feature is the
fact that global emission inventories track closer to methane emissions in the
most carbon-intensive scenario (RCP8.5) and that all other RCP scenarios remain
below the inventories. This suggests the tremendous challenge of climate
mitigation that lies ahead, particularly if current trajectories need to change
to be consistent with pathways leading to lower levels of global warming (Fig.
2).
|
|
| 3.1.3.0 | Methane emissions from fossil fuel production and use |
| 3.1.3.0 | Methane emissions from fossil fuel production and use | | | | Most of the methane anthropogenic emissions related to fossil fuels come
from the exploitation, transportation, and usage of coal, oil and natural gas.
This geological and fossil type of emission (see natural source section) is
driven by human activity. Additional emissions reported in this category
include small industrial contributions such as production of chemicals and
metals, and fossil fuel fires. Spatial distribution of methane emissions from
fossil fuel is presented in Fig. 3 based on the mean gridded maps provided by
EDGARv4.2FT2010 and GAINS over the 2003–2012 decade.
Global emissions of methane from fossil fuels and other industries are
estimated from three global inventories in the range of 114–133 Tg CH4
yr−1 for
the 2003–2012 decade with an average of 121 Tg CH4 yr-1
(Table 2), but with a large difference in the rate of change depending on
inventories. It represents on average 34 % (range 32–39 %) of the total global
anthropogenic emissions.
|
|
| 3.1.3.1 | Coal mining |
| 3.1.3.1 | Coal mining | | | | During mining, methane is emitted from ventilation shafts, where large
volumes of air are pumped into the mine to keep methane at a rate below 0.5 %
to avoid accidental inflammation. To prevent the diffusion of methane in the
mining working atmosphere, boreholes are made in order to evacuate methane. In
countries of the Organization for Economic Cooperation and Development (OECD),
methane recuperated from ventilation shafts is used as fuel, but in many
countries it is still emitted into the atmosphere or flared, despite efforts
for coal-mine recovery under the UNFCCC Clean Development Mechanisms
(http://cdm.unfccc.int). Methane emissions also occur during post-mining
handling, processing, and transportation. Some CH4 is released from
coal waste piles and abandoned mines. Emissions from these sources are believed
to be low because much of the CH4 would likely be emitted within the
mine (IPCC, 2000).
Almost 40 % (IEA, 2012) of the world’s electricity is produced from coal.
This contribution grew in the 2000s at the rate of several per cent per year,
driven by Asian production where large reserves exist, but has stalled from
2011 to 2012. In 2012, the top 10 largest coal producing nations accounted for
88 % of total world emissions for coal mining. Among them, the top three
producers (China, USA and India) produced two-thirds of the total (CIA, 2016).
Global estimates of methane emissions from coal mining show a large variation,
in part due to the lack of comprehensive data from all major producing
countries. The range of coal mining emissions is estimated at 18–46 Tg of
methane for the year 2005, the highest value being from EDGARv4.2FT2010 and the
lower from USEPA.
As announced in Sect. 3.1.2, coal mining is the main source explaining the
differences observed between inventories at global scale (Fig. 2). Indeed, such
differences are explained mainly by the different CH4 emission
factors used for calculating the fugitive emissions of the coal mining in
China. Coal mining emission factors depend strongly on the type of coal
extraction (underground mining emitting up to 10 times more than surface
mining), the geological underground structure (very region-specific) and history
(basin uplift), and the quality of the coal (brown coal emitting more than hard
coal). The EDGARv4.2FT2012 seems to have overestimated by a factor of 2 the
emission factor for the coal mining in China and allocated this to very few
coal mine locations (hotspot emissions). A recent county-based inventory of
Chinese methane emissions also confirms the overestimate of about >+38 %
with total anthropogenic emissions estimated at 43 ± 6 Tg
CH4 yr-1 (Peng et al., 2016). Also, assimilating also 13CH4
data, Thompson et al. (2015)
showed that their prior (based on EDGARv4.2FT2010) overestimated the Chinese
methane emissions by 30 %; however, they found no significant difference in the
coal sector estimates between prior and posterior. EDGARv4.2 follows the IPCC guidelines
2006, which recommends region-specific data. However, the EDGARv4.2 inventory
compilation used the European averaged emission factor for CH4 from
coal mine production in substitution for missing data, which seems to be twice
too high in China. This highlights that
significant errors on emission estimates may result from inappropriate use of
some emission factor and that applying “Tier 1” for coal mine emissions is not
accurate enough, as stated by the IPCC guidelines. The upcoming new version of
EDGARv4.3.2 will revise this down and distribute the fugitive CH4
from coal mining to more than 80 times more coal mining locations in China.
For the 2003–2012 decade, methane emissions from coal mining are estimated
at 34 % of total fossil-fuel-related emissions of methane (41 Tg CH4
yr-1, range of 26–50), consistent with the AMAP report when
considering the evolution since 2005. An additional very small source
corresponds to fossil fuel fires (mostly underground coal fires, ∼
0.1 Tg yr-1, EDGARv4.2FT2010).
|
|
| 3.1.3.2 | Oil and natural gas systems |
| 3.1.3.2 | Oil and natural gas systems | | | | Natural gas is comprised primarily of methane, so any leaks during drilling
of the wells, extraction, transportation, storage, gas distribution, and
incomplete combustion of gas flares contribute to methane emissions (Lamb et
al., 2015; Shorter et al., 1996). Fugitive permanent emissions (e.g. due to
leaky valves and compressors) should be distinguished from intermittent
emissions due to maintenance (e.g. purging and draining of pipes). During
transportation, leakage can occur in gas transmission pipelines, due to
corrosion, manufacturing, welding, etc. According to Lelieveld et al. (2005),
the CH4 leakage from gas pipelines should be relatively low;
however, distribution networks in older cities have increased leakage, especially
those with cast-iron and unprotected steel pipelines. Recent measurement
campaigns in different cities in the USA and Europe also revealed that
significant leaks occur in specific locations (e.g. storage facilities, city
gates, well and pipeline pressurization/depressurization points) along the
distribution networks to the end-users (Jackson et al., 2014a; McKain et al.,
2015). However, methane emissions can vary a lot from one city to another
depending in part on the age of city infrastructure (i.e. older cities on
average have higher emissions). Ground movements (landslides, earthquakes,
tectonic movements) can also release methane. Finally, additional methane
emissions from the oil industry (e.g. refining) and production of charcoal are
estimated to be a few Tg CH4 yr-1 only (EDGARv4.2, 2011).
In many facilities, such as gas and oil fields, refineries and offshore
platforms, venting of natural gas is now replaced by flaring with a partial
conversion into CO2; these two processes are usually considered
together in inventories of oil and gas industries.
Methane emissions from oil and natural gas systems also vary greatly in
different global inventories (46 to 98 Tg yr-1 in 2005;
Höglund-Isaksson et al., 2015). The inventories rely on the same sources and
magnitudes regarding the activity data. Thus, the derived differences result
from different methodologies and parameters used, including both emission and
activity factors. Those factors are country or even site-specific, and the few
field measurements available often combine oil and gas activities (Brandt et
al., 2014) and remain largely unknown for most major oiland gas-producing
countries. Depending on the country, the emission factors reported may vary by
2 orders of magnitude for oil production and by 1 order of magnitude for gas
production (Table 5.5 of Höglund-Isaksson et al., 2015). The GAINS estimate of
methane emissions from oil production is 4 times higher than EDGARv4.2FT2010
and USEPA. For natural gas, the uncertainty is also large (factor of 2), albeit
smaller than for oil production. The difference in these estimates comes from
the methodology used. Indeed, during oil extraction, the gas generated can be
either recovered (re-injected or utilized as an energy source) or not recovered
(flared or vented to the atmosphere). The recovery rates vary from one country
to another (being much higher in the USA, Europe and Canada than elsewhere),
and accounting for country-specific rates of generation and recovery of
associated gas might lead to an amount of gas released into the atmosphere 4
times higher during oil production than when using default values
(Höglund-Isaksson, 2012). This difference in methodology explains, in part, why
GAINS estimates are higher than EDGARv4.2FT2010 and USEPA. Another challenge
lies in determining the amount of flared or vented unrecovered gas, with
venting emitting CH4, whereas flaring converts all or most methane
(often > 99 %) to CO2. The balance of flaring and venting also
depends on the type of oil: flaring is less common for heavy oil wells than
conventional ones (Höglund-Isaksson et al., 2015). Satellite images can detect
flaring (Elvidge et al., 2009, 2016) and may be used to verify the country
estimates, but such satellites cannot currently be used to estimate the
efficiency of CH4 conversion to CO2.
For the 2003–2012 decade, methane emissions from upstream and downstream
natural oil and gas sectors are estimated to represent about 65 % of total
fossil CH4 emissions (79 Tg CH4 yr-1, range of
69–88, Table 2), with a lower uncertainty range than for coal emissions for
most countries.
|
|
| 3.1.3.3 | Shale gas |
| 3.1.3.3 | Shale gas | | | | Production of natural gas from the exploitation of hitherto unproductive
rock formations, especially shale, began in the 1980s in the US on an
experimental or small-scale basis. Then, from early 2000s, exploitations
started at large commercial scale. Two techniques developed and often applied
together are horizontal drilling and hydraulic fracturing. The shale gas
contribution to total natural gas production in the United States reached 40 %
in 2012, growing rapidly from only small volumes produced before 2005 (EIA,
2015). Indeed, the practice of high-volume hydraulic fracturing (fracking) for
oil and gas extraction is a growing sector of methane and other hydrocarbon production,
especially in the US. Most recent studies (Miller et al., 2013; Moore et al.,
2014; Olivier and Janssens-Maenhout, 2014; Jackson et al., 2014b; Howarth et
al., 2011; Pétron et al., 2014; Karion et al., 2013) albeit not all (Allen et
al., 2013; Cathles et al., 2012; Peischl et al., 2015) suggest that methane
emissions are underestimated by inventories and agencies, including the USEPA.
For instance, emissions in the Barnett Shale region of Texas from both
bottom-up and top-down measurements showed that methane emissions from upstream
oil and gas infrastructure were 90 % larger than estimates based on the USEPA’s
inventory and corresponded to 1.5 % of natural gas production (Zavala-Araiza et
al., 2015). This study also showed that a few high emitters, neglected in the
inventories, dominated emissions. Moreover these high emitting points, located
on the conventional part of the facility, could be avoided through better
operating conditions and repair of malfunctions. It also suggests that emission
factor of conventional and non-conventional gas facilities might not be as
different as originally thought (Howarth et al., 2011). Field measurements
suggest that emission factors for unconventional gas are higher than for
conventional gas, though the uncertainty, largely site-dependent, is large,
ranging from small leakage rate of 1–2 % (Peischl et al., 2015) to widely
spread rates of 3–17 % (Caulton et al., 2014; Schneising et al., 2014). For
current technology, the GAINS model has adopted an emission factor of 4.3 % for
shale-gas mining, still awaiting a clear consensus across studies.
|
|
| 3.1.4.0 | Agriculture and waste |
| 3.1.4.0 | Agriculture and waste | | | | This category includes methane emissions related to livestock (enteric
fermentation and manure), rice cultivation, landfills, and waste-water handling.
Of all types of emission, livestock is by far the largest emitter of CH4,
followed by waste handling and rice cultivation. Field burning of agricultural
residues was a minor source of CH4 reported in emission inventories.
The spatial distribution of methane emissions from agriculture and waste
handling is presented in Fig. 3 based on the mean gridded maps provided by
EDGARv4.2FT2010 and GAINS over the 2003– 2012 decade.
Global emissions for agriculture and waste are estimated at 195 Tg CH4
yr-1 (range 178–206, Table 2), representing 57 % of total
anthropogenic emissions.
|
|
| 3.1.4.1 | Livestock: enteric fermentation and manure management |
| 3.1.4.1 | Livestock: enteric fermentation and manure management | | | | Domestic livestock such as cattle, buffalo, sheep, goats, and camels produce
a large amount of methane by anaerobic microbial activity in their digestive
systems (Johnson et al., 2002). A very stable temperature (39 ◦C), a
stable pH (6.5– 6.8) in their rumen, and constant flow of plants (cattle graze
many hours per day) induce a production of metabolic hydrogen, used by methanogenic
Archaea together with CO2
to produce methane. The methane and carbon dioxide are released from the rumen
mainly through the mouth of multi-stomached ruminants (eructation, ∼
87 % of emissions) or absorbed in the blood system. The methane produced in the
intestines and partially transmitted through the rectum is only
∼
13 %. There are about 1.4 billion cattle globally, 1 billion sheep, and nearly
as many goats. The total number of animals is growing steadily
(http://faostat3.fao.org), although the number is not linearly related to the
CH4 emissions they produce; emissions are strongly influenced by the
total weight of the animals and their diet. Cattle, due to their large
population, large size, and particular digestive characteristics, account for
the majority of enteric fermentation CH4 emissions from livestock,
particularly, in the United States (USEPA, 2016). Methane emissions from
enteric fermentation are also variable from one country to another as cattle experience
water-limited conditions that highly vary spatially and temporally (especially
in the tropics).
In addition, when livestock or poultry manure are stored or treated in
systems that promote anaerobic conditions (e.g. as a liquid/slurry in lagoons,
ponds, tanks, or pits), the decomposition of the volatile solids component in
the manure tends to produce CH4. When manure is handled as a solid
(e.g. in stacks or dry lots) or deposited on pasture, range, or paddock lands,
it tends to decompose aerobically and produce little or no CH4.
Ambient temperature, moisture, and manure storage or residency time affect the
amount of CH4 produced because they influence the growth of the
bacteria responsible for CH4 formation. For non-liquid-based manure
systems, moist conditions (which are a function of rainfall and humidity) can
promote CH4 production. Manure composition, which varies with animal
diet, growth rate, and type, including the animal’s digestive system, also
affects the amount of CH4 produced. In general, the greater the
energy contents of the feed, the greater the potential for CH4
emissions. However, some higher-energy feeds also are more digestible than
lower quality forages, which can result in less overall waste excreted from the
animal (USEPA, 2006).
In 2005, global methane emissions from enteric fermentation and manure are
estimated in the range of 96– 114 Tg CH4 yr-1 in the
GAINS model and USEPA inventory, respectively, and in the range of 98–105 Tg CH4
yr-1 suggested by Kirschke et al. (2013). They are consistent with
the FAO-CH4 estimate of 102 Tg CH4 yr-1 for
2005 (Tubiello et al., 2013).
Here, for the 2003–2012 decade, based on all the databases aforementioned, we infer a range of 97–111
Tg CH4 yr-1 for the combination of enteric fermentation
and manure with a mean value of 106 Tg CH4 yr-1 (Table
2), about one-third of total global anthropogenic emissions.
|
|
| 3.1.4.2 | Waste management |
| 3.1.4.2 | Waste management | | | | This sector includes emissions from managed and nonmanaged landfills (solid waste disposal on land), and wastewater handling, where all kinds of waste are deposited, which can emit significant amounts of methane by anaerobic decomposition of organic material by microorganisms. Methane production from waste depends on pH, moisture and temperature. The optimum pH for methane emission is between 6.8 and 7.4 (Thorneloe et al., 2000). The development of carboxylic acids leads to low pH, which limits methane emissions. Food or organic waste, leaves and grass clippings ferment quite easily, while wood and wood products generally ferment slowly, and cellulose and lignin even more slowly (USEPA, 2010b). Waste management is responsible for about 11 % of total global anthropogenic methane emissions in 2000 at global scale (Kirschke et al., 2013). A recent assessment of methane emissions in the US accounts landfills for almost 26 % of total US anthropogenic methane emissions in 2014, the largest contribution of any CH4 source in the United States (USEPA, 2016). In Europe, gas control is mandatory on all landfills from 2009 onwards, following the ambitious objective raised in the EU Landfill Directive (1999) to reduce the landfilling of biodegradable waste by 65 % below the 1990 level by 2016. This is attempted through source separation and treatment of separated biodegradable waste in composts, biodigesters and paper recycling. This approach is assumed more efficient in terms of reducing methane emissions than the more usual gas collection and capture. Collected biogas is either burned by flaring or used as fuel if it is pure enough (i.e. the content of methane is > 30 %). Many managed landfills have the practice to apply cover material (e.g. soil, clay, sand) over the waste being disposed of in the landfill to prevent odour, reduce risk to public health, as well as to promote microbial communities of methanotrophic organisms (Bogner et al., 2007). In developing countries, very large open landfills still exist, with important health and environmental issues in addition to methane emissions (André et al., 2014). Waste water from domestic and industrial sources is treated in municipal sewage treatment facilities and private effluent treatment plants. The principal factor in determining the CH4 generation potential of waste water is the amount of degradable organic material in the waste water. Waste water with high organic content is treated anaerobically and that leads to increased emissions (André et al., 2014). The large and fast urban development worldwide, and especially in Asia, could enhance methane emissions from waste if adequate policies are not designed and implemented rapidly. The inventories give robust emission estimates from solid waste in the range of 28–44 Tg CH4 yr-1 in the year 2005, and waste water in the range 9–30 Tg CH4 yr-1 given by GAINS model and EDGAR inventory. In this study, global emissions of methane from landfills and waste are estimated in the range of 52– 63 Tg CH4 yr-1 for the 2003–2012 period with a mean value of 59 Tg CH4 yr1 , about 18 % of total global anthropogenic emissions. Rice cultivation Most of the world’s rice is grown on flooded fields (Baicich, 2013). Under these shallow-flooded conditions, aerobic decomposition of organic matter gradually depletes most of the oxygen in the soil, resulting in anaerobic conditions under which methanogenic Archaea decompose organic matter and produce methane. Most of this methane is oxidized in the underlying soil, while some is dissolved in the floodwater and leached away. The remaining methane is released to the atmosphere, primarily by diffusive transport through the rice plants, but also methane escapes from the soil via diffusion and bubbling through floodwaters (USEPA, 2016; Bridgham et al., 2013). The water management systems used to cultivate rice are one of the most important factors influencing CH4 emissions and is one of the most promising approach to mitigate the CH4 emissions from rice cultivation (e.g. periodical drainage and aeration not only causes existing soil CH4 to oxidize but also inhibits further CH4 production in soils (Simpson et al., 1995; USEPA, 2016; Zhang et al., 2016). Upland rice fields are not flooded and, therefore, are not believed to produce much CH4. Other factors that influence CH4 emissions from flooded rice fields include fertilization practices (i.e. the use of urea and organic fertilizers), soil temperature, soil type (texture and aggregated size), rice variety and cultivation practices (e.g. tillage, seeding, and weeding practices) (USEPA, 2011, 2016; Kai et al., 2011; Yan et al., 2009; Conrad et al., 2000). For instance, methane emissions from rice paddies increase with organic amendments (Cai et al., 1997) but can be mitigated by applying other types of fertilizers (mineral, composts, biogas residues, wet seeding) (Wassmann et al., 2000). Some studies have suggested that decreases in microbial emissions, particularly due to changes in the practice of rice cultivation, could be responsible for a ~ 15 Tg CH4 yr-1 decrease over the period from 1980s to 2000s (Kai et al., 2011). The geographical distribution of the emissions is assessed by global (USEPA, 2006, 2012; EDGARv4.2FT2010, 2013) and regional (Peng et al., 2016; Chen et al., 2013; Chen and Prinn, 2006; Yan et al., 2009; Castelán-Ortega et al., 2014; Zhang et al., 2014) inventories or by land surface models (Spahni et al., 2011; Zhang and Chen, 2014; Ren et al., 2011; Tian et al., 2010, 2011; Li et al., 2005; Pathak et al., 2005). The emissions show a seasonal cycle, peaking in the summer months in the extratropics associated with the monsoon and land management. Similar to emissions from livestock, emissions from rice paddies are influenced not only by extent of rice field area (equivalent to the number of livestock) but also by changes in the productivity of plants as these alter the CH4 emission factor used in inventories. The largest emissions are found in Asia (Hayashida et al., 2013), with China (5–11 Tg CH4 yr-1 ; Chen et al., 2013; Zhang et al., 2016) and India (~ 3–5 Tg CH4 yr-1 ; Bhatia et al., 2013) accounting for 30 to 50 % of global emissions |
|
| 3.1.4.3 | Rice cultivation |
| 3.1.4.3 | Rice cultivation | | | | Most of the world’s rice is grown on flooded fields (Baicich, 2013). Under
these shallow-flooded conditions, aerobic decomposition of organic matter
gradually depletes most of the oxygen in the soil, resulting in anaerobic
conditions under which methanogenic Archaea decompose organic matter and
produce methane. Most of this methane is oxidized in the underlying soil, while
some is dissolved in the floodwater and leached away. The remaining methane is
released to the atmosphere, primarily by diffusive transport through the rice
plants, but also methane escapes from the soil via diffusion and bubbling
through floodwaters (USEPA, 2016; Bridgham et al., 2013).
The water management systems used to cultivate rice are one of the most
important factors influencing CH4 emissions and is one of the most promising
approach to mitigate the CH4 emissions from rice cultivation (e.g. periodical
drainage and aeration not only causes existing soil CH4 to oxidize but also
inhibits further CH4 production in soils (Simpson et al., 1995; USEPA, 2016;
Zhang et al., 2016). Upland rice fields are not flooded and, therefore, are not
believed to produce much CH4. Other factors that influence CH4 emissions from
flooded rice fields include fertilization practices (i.e. the use of urea and
organic fertilizers), soil temperature, soil type (texture and aggregated
size), rice variety and cultivation practices (e.g. tillage, seeding, and
weeding practices) (USEPA, 2011, 2016; Kai et al., 2011; Yan et al., 2009;
Conrad et al., 2000). For instance, methane emissions from rice paddies
increase with organic amendments (Cai et al., 1997) but can be mitigated by
applying other types of fertilizers (mineral, composts, biogas residues, wet
seeding) (Wassmann et al., 2000). Some studies have suggested that decreases in
microbial emissions, particularly due to changes in the practice of rice
cultivation, could be responsible for a ∼ 15 Tg CH4 yr−1
decrease over the period from 1980s to 2000s (Kai et al., 2011).
The geographical distribution of the emissions is assessed by global (USEPA,
2006, 2012; EDGARv4.2FT2010, 2013) and regional (Peng et al., 2016; Chen et
al., 2013; Chen and Prinn, 2006; Yan et al., 2009; Castelán-Ortega et al.,
2014; Zhang et al., 2014) inventories or by land surface models (Spahni et al.,
2011; Zhang and Chen, 2014; Ren et al., 2011; Tian et al., 2010, 2011; Li et
al., 2005; Pathak et al., 2005). The emissions show a seasonal cycle, peaking
in the summer months in the extratropics associated with the monsoon and land
management. Similar to emissions from livestock, emissions from rice paddies
are influenced not only by extent of rice field area (equivalent to the number
of livestock) but also by changes in the productivity of plants as these alter
the CH4 emission factor used in inventories.
The largest emissions are found in Asia (Hayashida et al., 2013), with China
(5–11 Tg CH4 yr-1; Chen et al., 2013; Zhang et al., 2016)
and India (∼ 3–5 Tg CH4 yr-1; Bhatia et
al., 2013) accounting for 30 to 50 % of global emissions (Fig. 3). The decrease
of CH4 emissions from rice cultivation over the past decades is
confirmed in most inventories, because of the decrease in rice cultivation
area, the change in agricultural practices, and a northward shift of rice
cultivation since 1970s (e.g. Chen et al., 2013). Furthermore, recent studies
revealed that, together, high carbon dioxide concentrations and warmer
temperatures predicted for the end of the twenty-first century will about
double the amount of methane emitted per kilogramme of rice produced (van
Groenigen et al., 2013).
Based on global inventories only, global methane emissions from rice paddies
are estimated in the range 24–36 Tg CH4 yr−1 for the
2003–2012 decade, with a mean value of 30 Tg CH4 yr-1
(Table 2), about 9 % of total global anthropogenic emissions. The lower
estimate (24 Tg CH4 yr-1) is provided by FAO-CH4
inventory (Tubiello et al., 2013), which is based on a mix of FAO statistics
for crop production and IPCC guidelines.
|
|
| 3.1.5.0 | Biomass and biofuel burning |
| 3.1.5.0 | Biomass and biofuel burning | | | | This category includes all the combustion processes: biomass (forests,
savannahs, grasslands, peats, agricultural residues) and biofuels in the
residential sector (stoves, boilers, fireplaces). Biomass and biofuel burning
emits methane under incomplete combustion conditions, when oxygen availability
is insufficient such as charcoal manufacture and smouldering fires. The amount
of methane that is emitted during the burning of biomass depends primarily on
the amount of biomass, the burning conditions, and the material being burned. At
the global scale, biomass and biofuel burning lead to methane emissions of
27–35 Tg CH4 yr-1 with an average of 30 Tg CH4
yr-1 (2003–2012 decade, Table 2), of which 30–50 % is biofuel
burning (Kirschke et al., 2013).
In this study, we use the large-scale biomass burning (forest, savannah,
grassland and peat fires) from specific biomass burning inventories and the
biofuel burning contribution for the inventories (USEPA, GAINS and EDGAR).
The spatial distribution of methane emissions from biomass burning over the
2003–2012 decade is presented in Fig. 3 and is based on the mean gridded maps
provided by EDGARv4.2FT2010 and GAINS for the biofuel burning, and based on the
mean gridded maps provided by the biomass burning inventories presented
thereafter.
|
|
| 3.1.5.1 | Biomass burning |
| 3.1.5.1 | Biomass burning | | | | This category includes all the combustion processes: biomass (forests, savannahs, grasslands, peats, agricultural residues) and biofuels in the residential sector (stoves, boilers, fireplaces). Biomass and biofuel burning emits methane under incomplete combustion conditions, when oxygen availability is insufficient such as charcoal manufacture and smouldering fires. The amount of methane that is emitted during the burning of biomass depends primarily on the amount of biomass, the burning conditions, and the material being burned. At the global scale, biomass and biofuel burning lead to methane emissions of 27–35 Tg CH4 yr-1 with an average of 30 Tg CH4 yr-1 (2003–2012 decade, Table 2), of which 30–50 % is biofuel burning (Kirschke et al., 2013). In this study, we use the large-scale biomass burning (forest, savannah, grassland and peat fires) from specific biomass burning inventories and the biofuel burning contribution for the inventories (USEPA, GAINS and EDGAR). The spatial distribution of methane emissions from biomass burning over the 2003–2012 decade is presented in Fig. 3 and is based on the mean gridded maps provided by EDGARv4.2FT2010 and GAINS for the biofuel burning, and based on the mean gridded maps provided by the biomass burning inventories presented thereafter. |
|
| 3.1.5.2 | Biofuel burning |
| 3.1.5.2 | Biofuel burning | | | | Biomass that is used to produce energy for domestic, industrial, commercial,
or transportation purposes is hereafter called biofuel burning. A largely
dominant fraction of methane emissions from biofuels comes from domestic
cooking or heating in stoves, boilers and fireplaces, mostly in open cooking
fires where wood, charcoal, agricultural residues or animal dung are burnt.
Although more than 2 billion people, mostly in developing and emerging
countries, use solid biofuels to cook and heat their homes on a daily basis
(André et al., 2014), methane emissions from biofuel combustion have not yet
received the attention it should have to estimate its magnitude. Other much
smaller contributors include agricultural burning (∼ 1–2
Tg yr-1) and road transportation (< 1 Tg yr-1). Biofuel
burning estimates are gathered from USEPA, GAINS and EDGAR inventories.
In this study, biofuel burning is estimated to contribute 12 Tg CH4
yr-1 [10–14] to the global methane budget, about 3 % of total global
anthropogenic emissions.
|
|
| 3.2.0.0 | Natural methane sources |
| 3.2.0.0 | Natural methane sources | | | | Natural methane sources include wetland emissions as well as emissions from
other land water systems (lakes, ponds, rivers, estuaries), land geological
sources (seeps, microseepage, mud volcanoes, geothermal zones, and volcanoes,
marine seepages), wild animals, wildfires, termites, terrestrial permafrost and
oceanic sources (geological and biogenic). Many sources have been recognized
but their magnitude and variability remain uncertain (USEPA, 2010a; Kirschke et
al., 2013).
|
|
| 3.2.1.0 | Wetlands |
| 3.2.1.0 | Wetlands | | | | Wetlands are generally defined as ecosystems in which water saturation or
inundation (permanent or not) dominates the soil development and determines the
ecosystem composition (USEPA, 2010a). Such a broad definition needs to be
refined when it comes to methane emissions. In this work, we define wetlands as
ecosystems with inundated or saturated soils where anaerobic conditions lead to
methane production (USEPA, 2010a; Matthews and Fung, 1987). This includes
peatlands (bogs and fens), mineral wetlands (swamps and marshes), and seasonal
or permanent floodplains. It excludes exposed water surfaces without emergent
macrophytes, such as lakes, rivers, estuaries, ponds, and dams (addressed in
the next section), as well as rice agriculture (see Sect. 3.1.4., rice cultivation
paragraph). Even with this definition, one can consider that part of the
wetlands could be considered as anthropogenic systems, being affected by
human-driven land-use changes (Woodward et al., 2012). In the following we keep
the generic denomination wetlands for natural and human-influenced wetlands.
A key feature of wetland systems producing methane is anaerobic soils, where
high water table or flooded conditions limit oxygen availability and create
conditions for methanogenesis. In anoxic conditions, organic matter can be
degraded by methanogens that produce CH4. The three most important
factors influencing methane production in wetlands are the level of anoxia
(linked to water table), temperature and substrate availability (Wania et al.,
2010; Valentine et al., 1994; Whalen, 2005). Once produced, methane can reach
the atmosphere through a combination of three processes: molecular diffusion,
plant-mediated transport and ebullition. On its way to the atmosphere, methane
can be partly or completely oxidized by a group of bacteria, called
methanotrophs, which use methane as their only source of energy and carbon
(USEPA, 2010a). Concurrently, methane from the atmosphere can diffuse into the
soil column and be oxidized (see Sect. 3.3.4).
Land surface models estimate CH4 emissions through a series of
processes, including CH4 production, CH4 oxidation and
transport and are further regulated by the changing environmental factors (Tian
et al., 2010; Xu et al., 2010; Melton et al., 2013). In these models, methane
emissions from wetlands to the atmosphere are computed as the product of an
emission density (which can be negative; mass per unit area and unit time)
multiplied by a wetland extent (see the model intercomparison studies by Melton
et al., 2013, and Bohn et al., 2015). The CH4 emission density is
represented in land surface models with varying levels of complexity. Many
models link CH4 emission to net primary production (NPP) though
production of exudates or litter and soil carbon to yield heterotrophic respiration
estimates. A proportion of the heterotrophic respiration estimate is then taken
to be CH4 production (Melton et al., 2013). The oxidation of
produced (and becoming atmospheric) methane in the soil column is then either
represented explicitly (e.g. Riley et al., 2011; Grant and Roulet, 2002), or
just fixed proportionally to the production (Wania et al., 2013).
In land surface models, wetland extent is either prescribed (from
inventories or remote-sensing data) or computed using hydrological models
accounting for the fraction of grid cell with flat topography prone to high
water table (e.g. Stocker et al., 2014; Kleinen et al., 2012), or from data
assimilation against remote-sensed observations (Riley et al., 2011). Mixed
approaches can also be implemented with tropical extent prescribed from remote
sensing and northern peatland extent explicitly computed (Melton et al., 2013).
Wetland extent appears to be a large contributor to uncertainties in methane
emissions from wetlands (Bohn et al., 2015). For instance, the maximum wetland
extent on a yearly basis appeared to be very different among land surface
models in Melton et al. (2013), ranging from 7 to 27 Mkm2.Passive and active remote-sensing data in the
microwave domain have been used to retrieve inundated areas, as with the Global
Inundation Extent from Multi-Satellites product (GIEMS, Prigent et al., 2007;
Papa et al., 2010). These remote-sensed data do not exactly correspond to
wetlands, as all flooded areas are not wetlands (in methane emission sense) and
some wetlands (e.g. northern bogs) are not always flooded. Inundated areas also
include inland water bodies (lakes, ponds, estuaries) and rice paddies, which
have to be filtered out to compute wetland emissions. Overall, current remote
sensing of wetlands tends to underestimate wetland extent partly because of
signal deterioration over dense vegetation and partly because microwave signals
only detect water above or at the soil surface and therefore do not detect
emitting peatlands that are not inundated (Prigent et al., 2007). For example,
the Global Lakes and Wetlands Dataset (GLWD) (Lehner and Döll, 2004) estimates
between 8.2 and 10.1 Mkm2 of wetlands globally, while remote-sensing inundation area is
smaller, i.e. ∼ 6 Mkm2(Prigent
et al., 2007). Some ancillary data used in the GIEMS processing are not
available after 2007 and have prevented so far the extension of the dataset
after 2007.
Integrated at the global scale, wetlands are the largest and most uncertain
source of methane to the atmosphere (Kirschke et al., 2013). An ensemble of
land surface models estimated the range of methane emissions of natural
wetlands
at 141–264 Tg CH4 yr-1 for the 1993–2004 period, with
a mean and 1σ value of 190 ±
39 Tg CH4 yr-1 (Melton et al., 2013). Kirschke et
al. (2013) assessed a consistently large emission range of 142–287 Tg CH4
yr-1, using the Melton et al. (2013) land surface models and
atmospheric inversions. These emissions represent about 30 % of the total
methane source. The large range in the estimates of wetland emissions results
from difficulties in defining wetland CH4producing areas as well as
in parameterizing terrestrial anaerobic sources and oxidative sinks (Melton et
al., 2013; Wania et al., 2013).
In this work, following Melton et al. (2013), 11 land surface models (Table
1) computing net CH4 emissions have been run under a common protocol
with a 30-year spin-up (1901–1930) followed by a simulation until the end of
2012 forced by CRU-NCEP v4.0 reconstructed climate fields. Atmospheric CO2 influencing
NPP was also prescribed in the models, allowing the models to separately
estimate carbon availability for methanogenesis. In all models, the same
wetland extent (SWAMPS-GLWD) has been prescribed. The SWAMPS-GLWD is a monthly
global wetland area dataset, which has been developed to overcome the
aforementioned issues and combines remote-sensing data from Schroeder et al.
(2015) and GLWD inventory in order to develop a monthly global wetland area
dataset (Poulter et al., 2016). Briefly, GLWD was used to set the annual mean
wetland area, to which a seasonal cycle of fractional surface water was added
using data from the Surface WAter Microwave Product Series Version 2.0 (SWAMPS)
(Schroeder et al., 2015). The combined GLWD-SWAMPS product leads to a maximum
annual wetland area of 10.5 Mkm2 (8.7 Mkm2on average, about 5.5 % of than global land
surface). The largest wetland areas in the SWAMPS-GLWD are in Amazonia, the
Congo Basin, and the western Siberian lowlands, which in previous studies have
appeared to be strongly underestimated by several inventories (Bohn et al.,
2015). However, wetlands above 70◦ N appear under-represented in GLWD as
compared to Sheng et al. (2004) and Peregon et al. (2008). Indeed,
approximately half of the global natural wetland area lies in the boreal zone
between 50 and 70◦ N, while 35 % can be found in the tropics, between 20◦
N and 30◦ S (Matthews and Fung, 1987; Aselmann and Crutzen, 1989).
Despite the lower area extent, the higher per-unit area methane emissions of
tropical wetlands results in a larger wetland source from the tropics than from
the boreal zone (Melton et al., 2013).
The average emission map from wetlands for 2003–2012 built from the 11
models is plotted in Fig. 3. The zones with the largest emissions reflect the
GLWD database: the Amazon basin, equatorial Africa and Asia, Canada, western
Siberia, eastern India, and Bangladesh. Regions where methane emissions are
robustly inferred (i.e. regions where mean flux is larger than the standard
deviation of the models) represent 80 % of the total methane flux due to
natural wetlands. Main primary emission zones are consistent between models,
which is clearly favoured by the common wetland extend prescribed. But still,
the different sensitivity of the models to temperature can generate substantial
different patterns, such as in India. Some secondary (in magnitude) emission
zones are also consistently inferred between models: Scandinavia, continental
Europe, eastern Siberia, central USA, and tropical Africa. Using improved regional
methane emission datasets (such as studies over North America, Africa, China,
and Amazon) can enhance the accuracy of the global budget assessment (Tian et
al., 2011; Xu and Tian, 2012; Ringeval et al., 2014; Valentini et al., 2014).
The resulting global flux range for natural wetland emissions is 153–227 Tg
CH4 yr-1 for the 2003–2012 decade, with an average of 185
Tg CH4 yr-1 with a 1σ
standard deviation of 21 Tg CH4 yr-1 (Table 2).
|
|
| 3.2.2.0 | Other inland water systems (lakes, ponds, rivers, estuaries) |
| 3.2.2.0 | Other inland water systems (lakes, ponds, rivers, estuaries) | | | | This category includes methane emissions from freshwater systems (lakes,
ponds, rivers) and from brackish waters of estuaries. Methane emissions from
fresh waters and estuaries occur through a number of pathways including (1)
continuous or episodic diffusive flux across water surfaces, (2) ebullition
flux from sediments, (3) flux mediated through the aerenchyma of emergent
aquatic macrophytes (plant transport) in littoral environments, and also for
reservoirs, (4) degassing of CH4 in the turbines, and (5) elevated
diffusive emissions in rivers downstream of the turbines especially if water
through the turbines is supplied from anoxic CH4-rich water layers
in the reservoir (Bastviken et al., 2004; Guérin et al., 2006, 2016). It is
very rare that complete emission budgets include all these types of fluxes. For
methodological reasons many past and present flux measurements only account for
the diffusive flux based on short-term flux chamber measurements where
non-linear fluxes were often discarded. At the same time, diffusive flux is now
recognized as a relatively small flux component in many lakes, compared to
ebullition and plant fluxes (in lakes with substantial emergent macrophyte
communities). The two latter fluxes are very challenging to measure, both
typically being associated with shallow near-shore waters and having high
spatiotemporal variability. Ebullition can also occur more frequently in areas
with high sediment organic matter load and is by nature episodic with very high
fluxes occurring over time frames of seconds followed by long periods without
ebullition.
Freshwater contributions from lakes were first estimated to emit 1–20 Tg CH4
yr-1 based on measurements in two systems (Great Fresh Creek,
Maryland, and Lake Erie; Ehhalt, 1974). A subsequent global emission estimate
was 11– 55 Tg CH4 yr-1 based on measurements from three
arctic lakes and a few temperate and tropical systems (Smith and Lewis, 1992),
and 8–48 Tg CH4 yr-1 using extended data from all of the
lake rich biomes (73 lakes; Bastviken et al., 2004). Combining results from
Bastviken et al. (2004) and Bastviken et al. (2011), Kirschke et al. (2013)
reported a range of 8–73 Tg CH4 yr-1. Gradually, methane
emissions from reservoirs and rivers have also been included in the most recent
global estimate from fresh waters of 103 Tg CH4 yr-1,
including emissions from non-saline lakes, reservoirs, ponds and rivers (data
from 473 systems; Bastviken et al., 2011). Improved stream and river emission
estimates of 27 Tg CH4 yr-1 were recently suggested
(Stanley et al., 2016). Importantly, the previous estimates of inland water
fluxes are not independent. Instead they represent updates from increasing data
quantity and quality. It should also be noted that issues regarding spatiotemporal
variability are not considered in consistent ways at present (Wik et al.,
2016a; Natchimuthu et al., 2015).
Present data do not allow for separating inland water fluxes over the
different time periods investigated in this paper. The global estimates provided
are therefore assumed to be constant for this study. Here we combine the latest
estimates of global freshwater CH4 emissions (Bastviken et al.,
2011) with a more recent regional estimate for latitudes above 50◦ N at
present (Wik et al., 2016b) and new extrapolations for tropical river emissions
(Borges et al., 2015; Sawakuchi et al., 2014) and streams (Stanley et al.,
2016). High-latitude lakes include both post-glacial lakes and thermokarst
lakes (water bodies formed by thermokarst), the latter having larger emissions
per square metre but smaller regional emissions than the former because of
smaller areal extent (Wik et al., 2016b). Water body depth, sediment type, and
ecoclimatic region are the key factors explaining variation in methane fluxes
from lakes (Wik et al., 2016b).
Altogether, these studies consider data from more than 900 systems, of which
∼
750 are located north of 50◦ N. In this context we only consider fluxes
from open waters assuming that plant-mediated fluxes are included in the wetland
emission term. The average total estimated open water emission including the
recent estimates from smaller streams is 122 Tg CH4 yr-1.
The uncertainty is high with a coefficient of variation ranging from 50 to >
100 % for various flux components and biomes (Bastviken et al., 2011) resulting
in a minimum uncertainty range of 60–180 Tg CH4 yr-1. The
present data indicate that lakes or natural ponds, reservoirs, and
streams/rivers account for 62, 16 and 22 % of the average fluxes, respectively
(given the large uncertainty the percentages should be seen as approximate
relative magnitudes only).
Potentially, the emissions from reservoirs should be allocated to
anthropogenic emissions (not done here). Regarding lakes and reservoirs,
tropical (< 30◦ latitude) and temperate (30–50◦ latitude)
emissions represent 49 and 33 % of the flux, respectively, with 18 % left for
regions above 50◦ latitude. For comparison, approximately 40 % of the
inland water surface area is found above 50◦ latitude in the Northern
Hemisphere and 34 % of the area is situated between 20◦ S and 20◦ N
(Verpoorter et al., 2014). Ebullition typically accounted for 50 to more than
90 % of the flux from the water bodies, while contributions from ebullition
appear lower from rivers, although this is currently debated (e.g. Crawford et
al., 2014). Several aspects will need consideration to reduce the remaining
uncertainty in the freshwater fluxes, including the generation of flux
measurement that is more representative in time and space and an update of
global lake area databases (e.g. GLOWAB, Verpoorter et al., 2014).
|
|
| 3.2.3.0 | Onshore and offshore geological sources |
| 3.2.3.0 | Onshore and offshore geological sources | | | | Significant amounts of methane, produced within the Earth’s crust, naturally
migrate to the atmosphere through tectonic faults and fractured rocks. Major
emissions are related to hydrocarbon production in sedimentary basins
(microbial and thermogenic methane), through continuous exhalation and
eruptions from onshore and shallow marine gas/oil seeps and through diffuse
soil microseepage (after Etiope, 2015). Specifically, six source categories
have been considered. Five are onshore sources: mud volcanoes (sedimentary
volcanism), gas and oil seeps (independent of mud volcanism), microseepage
(diffuse exhalation from soil in petroleum basins), geothermal (non-volcanic)
manifestations and volcanoes. One source is offshore: submarine seepage
(several types of gas manifestation at the seabed). Figure 4a shows the areas
and locations potentially emitting geological methane, showing diffuse
potential microseepage regions, macroseepage locations (oil–gas seeps, mud
volcanoes) and geothermal/volcanic areas (built from Etiope, 2015), which
represent more than 1000 emitting spots.
Studies since 2000 have shown that the natural release to the Earth’s
surface of methane of geological origin is an important global greenhouse gas
source (Etiope and Klusman, 2002; Kvenvolden and Rogers, 2005; Etiope et al.,
2008; USEPA, 2010a; Etiope, 2012, 2015). Indeed, the geological source is in
the top-three natural methane sources after wetlands (and with freshwater
systems) and about 10 % of total methane emissions, of the same magnitude or
exceeding other sources or sinks, such as biomass burning, termites and soil
uptake, considered in recent IPCC assessment reports (Ciais et al., 2013).
In this study, the following provided estimates were derived by bottom-up
approaches based on (a) the acquisition of thousands of land-based flux
measurements for various seepage types in many countries, and (b) the
application of the same procedures typically used for natural and anthropogenic
gas sources, following upscaling methods based on the concepts of “point
sources”, “area sources”, “activity” and “emission factors”, as recommended by
the air pollutant emission guidebook of the European Environment Agency
(EMEP/EEA, 2009). Our estimate is consistent with a top-down global
verification, based on observations of radiocarbon-free (fossil) methane in the
atmosphere (Etiope et al., 2008; Lassey et al., 2007b), with a range of 33–75
Tg CH4 yr-1.
As a result, in this study, the global geological methane emission is
estimated in the range of 35–76 Tg CH4 yr-1 (mean of 52
Tg CH4 yr-1), with 40 Tg CH4 yr−1 [30–56] for
onshore emissions (10–20 Tg CH4 yr-1 for mud volcanoes,
3–4 Tg yr-1 for gas–oil seeps, 10–25 Tg yr-1 for
microseepage, 2–7 Tg CH4 yr-1 for geothermal/volcanic
manifestations) and 12 Tg CH4 yr-1 [5–20] for offshore
emissions through marine seepage (Rhee et al., 2009; Berchet et al., 2016;
Etiope, 2012; see Sect. 3.2.6 for offshore contribution explanations).
|
|
| 3.2.4.0 | Termites |
| 3.2.4.0 | Termites | | | | Termites are important decomposer organisms, which play a very relevant role
in the cycling of nutrients in tropical and subtropical ecosystems (Sanderson,
1996). The degradation of organic matter in their gut, by symbiotic anaerobic
microorganisms, leads to the production of CH4 and CO2 (Sanderson,
1996). The upscaling approaches which have been used to quantify the
contribution of termites to global CH4 emissions (Sanderson, 1996;
Sugimoto et al., 1998; Bignell et al., 1997) are affected by large
uncertainties, mainly related to the effect of soil and mound environments on
net CH4 emissions; the quantification of termite biomass for each
ecosystem type; and the impact of land-use change on termite biomass. For all these
factors, uncertainty mainly comes from the relatively small number of studies
compared to other CH4 sources. In Kirschke et al. (2013) (see their
Supplement), a reanalysis of CH4 emissions from termites at the
global scale was proposed and CH4 emissions per unit of surface were
estimated as the product of termite biomass, termite CH4 emissions
per unit of termite mass and a scalar factor expressing the effect of
land-use/land-cover change. The latter two terms were estimated from published
literature reanalysis (Kirschke et al., 2013, Supplement). A climate zoning
(following the Köppen–Geiger classification) was applied to updated climate
datasets by Santini and Di Paola (2015) and was adopted to take into account
different combinations of termite biomass per unit area and CH4
emission factor per unit of termite biomass. In the case of tropical climate,
first termites’ biomass was estimated by a simple regression model representing
its dependence on gross primary productivity (Kirschke et al., 2013, Supplement),
whereas termites’ biomass for forest and grassland ecosystems of the warm
temperate climate and for shrublands of the Mediterranean subclimate were
estimated from data reported by Sanderson (1996). CH4 emission
factor per unit of termite biomass was derived from published literature and
was estimated equal to 2.8 mg CH4 g-1 termite h-1
for tropical ecosystems and Mediterranean shrublands (Kirschke et al., 2013)
and 1.7 mg CH4 g-1 termite h-1 for temperate
forests and grasslands (Fraser et al., 1986). Emissions were scaled up in GIS environment and annual CH4 fluxes
computed for the three periods 1982–1989, 1990–1999 and 2000–2007
representative of the 1980s, 1990s and 2000s, respectively. CH4
emissions showed only little interannual and interdecadal variability (0.1 Tg
CH4 yr-1) and strong regional variability with tropical
South America and Africa being the main sources (36 and 30 % of the global
total emissions, respectively) due to the extent of their natural forest and
savannah ecosystems (Fig. 4b). For the 2000s, a global total of 8.7 ±
3.1 Tg CH4 yr-1 (range 3–15 Tg CH4 yr-1)
was obtained. This value is close to the average estimate derived from previous
upscaling studies which report values spanning from 2 to 22 Tg CH4
yr-1 (Ciais et al., 2013). In this study, we adopt a value of 9 Tg
CH4 yr−1 (range 3–15 Tg CH4 yr-1, Table 2).
|
|
| 3.2.5.0 | Wild animals |
| 3.2.5.0 | Wild animals | | | | As for domestic ruminants, wild ruminants eruct or exhale methane through
the microbial fermentation process occurring in their rumen (USEPA, 2010a).
Global emissions of CH4 from wild animals range from 2–6 Tg CH4
yr-1 (Leng, 1993) to 15 Tg CH4 yr-1 (Houweling
et al., 2000). The global distribution of CH4 emissions from wild
ruminants is generally estimated as a function of the percentage and type of
vegetation consumed by the animals (Bouwman et al., 1997). However, as
suspected, numerous and various wild animals live partly hidden in the forests,
savannahs, etc., challenging the assessment of these emissions.
The range adopted in this study is 2–15 Tg CH4 yr-1
with a mean value of 10 Tg CH4 yr-1 (Table 2).
|
|
| 3.2.6.0 | Oceanic sources |
| 3.2.6.0 | Oceanic sources | | | | Possible sources of oceanic CH4 include the following: (1) leaks
from geological marine seepage (see also Sect. 3.2.3); (2) production from
sediments or thawing subsea permafrost; (3) emission from the destabilization
of marine hydrates and (4) in situ production in the water column, especially
in the coastal ocean because of submarine groundwater discharge (USEPA, 2010a).
Once at seabed, methane can be transported through the water column by
diffusion in a dissolved form (especially in the upwelling zones) or by
ebullition (gas bubbles, e.g. from geological marine seeps), for instance, in
shallow waters of continental shelves. Among these different origins of oceanic
methane, hydrates have attracted a lot of attention. Methane hydrates (or
sometimes called clathrates) are ice-like crystals formed under specific
temperature and temperature conditions (Milkov, 2005). The stability zone for
methane hydrates (high pressure, ambient temperatures) can be found in the
shallow lithosphere (i.e. < 2000 m depth), either in the continental
sedimentary rocks of polar regions or in the oceanic sediments at water depths
greater than 300 m (continental shelves, sediment– water interface) (Kvenvolden
and Rogers, 2005; Milkov, 2005). Methane hydrates can be either of biogenic
origin (formed in situ at depth in the sediment by microbial activity) or of
thermogenic origin (non-biogenic gas migrated from deeper sediments and trapped
due to pressure/temperature conditions or due to some capping geological
structure such as marine permafrost). The total stock of marine methane
hydrates is large but uncertain, with global estimates ranging from hundreds to
thousands of Pg CH4 (Klauda and Sandler, 2005; Wallmann et al.,
2012).
If the production of methane at seabed can be of importance, for instance,
marine seepages emit up to 65 Tg CH4 yr-1 globally at
seabed level (USEPA, 2010a); more uncertain is the flux of oceanic methane
reaching the atmosphere. For example, bubble plumes of CH4 from the
seabed have been observed in the water column but not detected in the Arctic
atmosphere (Westbrook et al., 2009; Fisher et al., 2011). A large part of the
seabed CH4 production and emission is oxidized in the water column
and does not reach the atmosphere (James et al., 2016). There are several
barriers preventing methane from being expelled to the atmosphere. From the
bottom to the top, gas hydrates and permafrost serve as a barrier to fluid and
gas migration towards the seafloor (James et al., 2016). First, on centennial
to millennium timescales, trapped gases may be released when permafrost is
perturbed and cracks or through Pingolike features. At present, microbial
processes are the most important control on methane emissions from marine
environments. Aerobic oxidation in the water column is a very efficient sink,
which allows very little methane even from established and vigorous gas seep
areas or even gas well blowouts such as the Deepwater Horizon from reaching the
atmosphere. Anaerobic methane oxidation, first described by Reeburgh and Heggie
(1977), coupled to sulfate reduction controls methane losses from sediments to
the overlying water (Reeburgh, 2007). Methane only escapes marine sediments in
significant amounts from rapidly accumulating sedimentary environments or via
advective processes such as ebullition or groundwater flow in shallow shelf
regions. Anaerobic methane oxidation was recently demonstrated to be able to
keep up with the thaw front of thawing permafrost in a region that had been
inundated within the past 1000 years (Overduin et al., 2015). Second, the
oceanic pycnocline is a physical barrier limiting the transport of methane (and
other species) towards the surface. Third, another important mechanism stopping
methane from reaching the ocean surface is the dissolution of bubbles into the
ocean water. Although bubbling is the most efficient way to transfer methane
from the seabed to the atmosphere, the fraction of bubbles actually reaching
the atmosphere is very uncertain and critically depends on emission depths
(< 100–200 m, McGinnis et al., 2015) and on the size of the bubbles (>
5–8 mm; James et al., 2016). Finally, surface oceans are aerobic and contribute
to the oxidation of dissolved methane (USEPA, 2010a). However, surface waters
can be more supersaturated than the underlying deeper waters, leading to a
methane paradox (Sasakawa et al., 2008). Possible explanations involve
upwelling in areas with surface mixed layers covered by sea ice (Damm et al.,
2015) or methane produced within the anoxic centre of sinking particles
(Sasakawa et al., 2008), but more work is needed to correct such an apparent
paradox.
All published estimates agree that contemporary global methane emissions
from oceanic sources are only a small contributor to the global methane budget,
but the range of estimates is relatively large from 1 to 35 Tg CH4
yr-1 when summing geological and other emissions (e.g. Rhee et al.,
2009; Etiope, 2015; USEPA, 2010a). For geological emissions, the most used
value is 20 Tg yr-1, relying on expert knowledge and literature
synthesis proposed in a workshop reported in Kvenvolden et al. (2001); the
authors of this study recognized that this first estimation needs to be
revised. Since then, oceanographic campaigns have been organized, especially to
sample bubbling areas. For instance, Shakhova et al. (2010, 2014) infer 8–17 Tg
CH4 yr-1 emissions just for the East Siberian Arctic
Shelf (ESAS), based on the extrapolation of numerous but local measurements,
and possibly related to melting seabed permafrost (Shakhova et al., 2015).
Because of the highly heterogeneous distribution of dissolved CH4 in
coastal regions, where bubbles can reach the atmosphere, extrapolation of in
situ local measurements to the global scale can be hazardous and lead to biased
global estimates. Indeed, using very precise and accurate continuous
atmospheric methane observations in the Arctic region, Berchet et al. (2016)
showed that Shakhova’s estimates are 4–8 times too large to be compatible with
atmospheric signals. This recent result suggests that the current estimate of
20 Tg yr-1 for the global emissions due to geological seeps
emissions to the atmosphere in coastal oceans is too large and needs revision.
Applying crudely the Berchet et al. (2016) abatement factor leads to emissions
as low as less than 5 Tg CH4 yr-1.
More studies are needed to sort out this discrepancy and we choose to report
here the full range of 5–20 Tg CH4 yr-1 for marine
geological emissions, with a mean value of 12 Tg CH4 yr-1.
Concerning non-geological ocean emissions (biogenic, hydrates), the most
common value found in the literature is 10 Tg CH4 yr-1
(Rhee et al., 2009). It appears that most studies rely on the work of Ehhalt
(1974), where the value was estimated on the basis of the measurements done by
Swinnerton and co-workers (Lamontagne et al., 1973; Swinnerton and Linnenbom,
1967) for the open ocean, combined with
purely speculated emissions from the continental shelf. Based on basin-wide
observations using updated methodologies, three studies found estimates ranging
from 0.2 to 3 Tg CH4 yr-1 (Conrad and Seiler, 1988; Bates
et al., 1996; Rhee et al., 2009), associated with supersaturations of surface
waters that are an order of magnitude smaller than previously estimated, both
for the open ocean (saturation anomaly ∼ 0.04, see Rhee et al.,
2009, Eq. 4) and for the continental shelf (saturation anomaly ∼
0.2). In their synthesis indirectly referring to the original observations from
Lambert and Schmidt (1993), Wuebbles and Hayhoe (2002) use a value of 5 Tg CH4
yr-1. Proposed explanations for discrepancies regarding sea-to-air
methane emissions in the open ocean rely on experimental biases in the former
study of Swinnerton and Linnenbom (1967) (Rhee et al., 2009). This may explain
why the Bange et al. (1994) compilation cites a global source of 11–18 Tg CH4
yr-1 with a dominant contribution of coastal regions. Here, we
report a range of 0– 5 Tg CH4 yr-1, with a mean value of
2 Tg CH4 yr-1.
Concerning more specifically atmospheric emissions from marine hydrates,
Etiope (2015) points that current estimates of methane air–sea flux from
hydrates (2–10 Tg CH4 yr-1 in e.g. Ciais et al., 2013, or
Kirschke et al., 2013) originate from the hypothetical values of Cicerone and
Oremland (1988). No experimental data or estimation procedures have been
explicitly described along the chain of references since then (Lelieveld et
al., 1998; Denman et al., 2007; Kirschke et al., 2013; IPCC, 2001). It was
recently estimated that ∼ 473 Tg CH4 was
released in the water column over 100 years (Kretschmer et al., 2015). Those
few Tg per
year become negligible once consumption in the water column has been
accounted for. While events such as submarine slumps may trigger local releases
of considerable amounts of methane from hydrates that may reach the atmosphere
(Etiope, 2015; Paull et al., 2002), on a global scale, presentday atmospheric
methane emissions from hydrates do not appear to be a significant source to the
atmosphere.
Overall, these elements suggest the necessity to revise to a lower value the
current total oceanic methane source to the atmosphere. Summing biogenic,
geological and hydrate emissions from oceans leads to a total oceanic methane
emission of 14 Tg CH4 yr-1 (range 5–25). Refining this
estimate requires performing more in situ measurements of atmospheric and
surface water methane concentrations and of bubbling areas and would require
the development of process-based models for oceanic methane linking sediment
production and oxidation, transport and transformation in the water column and
atmospheric exchange (James et al., 2016).
|
|
| 3.2.7.0 | Terrestrial permafrost and hydrates |
| 3.2.7.0 | Terrestrial permafrost and hydrates | | | | Permafrost is defined as frozen soil, sediment, or rock having temperatures
at or below 0 ◦C for at least two consecutive years (ACIA, 2005; Arctic
Research Commission, 2003). The total extent of permafrost zones of the
Northern Hemisphere is about 15 % of the land surface, with values around 15
million square kilometres (Slater and Lawrence, 2013; Levavasseur et al., 2011;
Zhang et al., 1999). Where soil temperatures have passed the 0 ◦C mark,
thawing of the permafrost at its margins occurs, accompanied by a deepening of
the active layer (Anisimov and Reneva, 2006) and possible formation of
thermokarst lakes (Christensen et al., 2015). A total of 1035 ±
150 Pg of carbon can be found in the upper 3 m or permafrost regions, or
∼
1300 Pg of carbon (1100 to 1500) Pg C for all permafrost (Hugelius et al.,
2014; Tarnocai et al., 2009). The thawing permafrost can generate direct and
indirect methane emissions. Direct methane emissions rely on the release of the
methane contained in the thawing permafrost.
This flux to the atmosphere is small and estimated to be at maximum 1 Tg CH4
yr-1 at present (USEPA, 2010a). Indirect methane emissions are
probably more important. They rely on the following: (1) methanogenesis induced
when the organic matter contained in thawing permafrost is released; the
associated changes in land surface hydrology possibly enhancing methane
production (McCalley et al., 2014); and the formation of more thermokarst lakes
from erosion and soil collapsing. Such methane production is probably already
significant today and could be more important in the future associated with a
strong positive feedback to climate change. However, indirect methane emissions
from permafrost thawing are difficult to estimate at present, with no data yet
to refer to, and in any case they largely overlap with wetland and freshwater
emissions occurring above or around thawing areas.
Here, we choose to report here only the direct emission range of 0–1 Tg CH4
yr-1, keeping in mind that current wetland, thermokarst lakes and
other freshwater methane emissions already likely include a significant
indirect contribution originating from thawing permafrost. For the next
century, it has been recently estimated that 5–15 % of the terrestrial
permafrost carbon pool is vulnerable to release in the form of greenhouse
gases, corresponding to 130–160 Pg C. The likely progressive release in the
atmosphere of such an amount of carbon as carbon dioxide and methane will have
a significant impact on climate change trajectory (Schuur et al., 2015). The
underlying methane hydrates represent a substantial reservoir of methane,
estimated up to 530 000 Tg of CH4 (Ciais et al., 2013). Present and
future emissions related to this reservoir are very difficult to assess at the
moment and require more studies.
|
|
| 3.2.8.0 | Vegetation |
| 3.2.8.0 | Vegetation | | | | A series of recent studies define three distinct pathways for the production
and emission of methane by living vegetation. First, plants produce methane
through an abiotic photochemical process induced by stress (Keppler et al.,
2006). This pathway was criticized (e.g. Dueck et al., 2007; Nisbet et al.,
2009), and although numerous studies have since confirmed aerobic emissions
from plants and better resolved its physical drivers (Fraser et al., 2015), global
estimates still vary by 2 orders of magnitude (Liu et al., 2015) meaning any
potential implication for the global methane budget remains highly uncertain.
Second, plants act as “straws”, drawing methane produced by microbes in anoxic
soils (Rice et al., 2010; Cicerone and Shetter, 1981). Third, the stems of
living trees commonly provide an environment suitable for microbial
methanogenesis (Covey et al., 2012). Static chambers demonstrate locally
significant through-bark flux from both soil-based (Pangala et al., 2013,
2015), and tree-stem-based methanogens (Wang et al., 2016). These studies
indicate trees are a significant factor regulating ecosystem flux; however,
estimates of biogenic plant-mediated methane emissions at broad scales are
complicated by overlap with methane consumption in upland soil and production
in wetlands. Integrating plant-mediated emissions in the global methane budget
will require untangling these processes to better define the mechanisms,
spatio-temporal patterns, and magnitude of these pathways.
|
|
| 3.3.0.0 | Methane sinks and lifetime |
| 3.3.0.0 | Methane sinks and lifetime | | | | Methane is the most abundant reactive trace gas in the troposphere and its
reactivity is important to both tropospheric and stratospheric chemistry. The
main atmospheric sink of methane is its oxidation by the hydroxyl radical (OH),
mostly in the troposphere, which contributes about 90 % of the total methane
sink (Ehhalt, 1974). Other losses are by photochemistry in the stratosphere
(reactions with chlorine atoms, Cl, and atomic oxygen, O(1D)), by
oxidation in soils (Curry, 2007; Dutaur and Verchot, 2007), and by
photochemistry in the marine boundary layer (reaction with Cl; Allan et al.,
2007; Thornton et al., 2010). Uncertainties in the total sink of methane as
estimated by atmospheric chemistry models are of the order of 20–40 % (Kirschke
et al., 2013). It is much less (10–20 %) when using atmospheric proxy methods
(e.g. methyl chloroform, see below) as in atmospheric inversions (Kirschke et
al., 2013). Methane is a significant source of water vapour in the middle to
upper stratosphere and influences stratospheric ozone concentrations by
converting reactive chlorine to less reactive hydrochloric acid (HCl). In the
present release of the global methane budget, we essentially rely on the former
analysis of Kirschke et al. (2013) and IPCC AR5. Following the ACCMIP model
intercomparison (Lamarque et al., 2013), the ongoing Climate Chemistry Model
Initiative (CCMI) and the upcoming Aerosols Chemistry Modeling Intercomparison
Project (AerChemMIP) should allow obtaining updated estimates on methane
chemical sinks and lifetimes.
|
|
| 3.3.1.0 | OH oxidation |
| 3.3.1.0 | OH oxidation | | | | OH radicals are produced following the photolysis of ozone (O3)in the presence of water vapour. OH is
destroyed by reactions with CO, CH4, and non-methane volatile organic
compounds, but since OH exists in photochemical equilibrium with HO2,
the net effect of CH4 oxidation on the HOx budget
also depends on the level of NOx(Lelieveld et al., 2002) and other
competitive oxidants. Considering its very short lifetime (a few seconds,
Lelieveld et al., 2004), it is not possible to estimate global OH
concentrations directly from observations. Observations are generally carried
out within the boundary layer, while the global OH distribution and variability
are more influenced by the free troposphere (Lelieveld et al., 2016). A series
of experiments were conducted by several chemistryclimate models and chemistry
transport models participating in the Atmospheric Chemistry and Climate Model
Intercomparison Project (ACCMIP) to study the long-term changes in atmospheric
composition between 1850 and 2100 (Lamarque et al., 2013). For the year 2000,
the multimodel mean (14 models) global mass-weighted OH tropospheric
concentration is 11.7 ± 1.0 ×
105 molec
cm−3(range
10.3–13.4 ×
105 molec
cm−3, Voulgarakis et al., 2013), consistent with the
estimates of Prather et al. (2012) at 1.2 ± 1.3 ×
105 molec
cm−3. However, it is worth noting that, in the ACCMIP
estimations, the differences in global
OH are larger between models than between pre-industrial, present and future
emission scenario simulations. Indeed Lelieveld et al. (2016) suggest that
tropospheric OH is buffered against potential perturbations from emissions,
mostly due to chemistry and transport connections in the free troposphere,
through transport of oxidants such as ozone. Besides the uncertainty on global
OH concentrations, the OH distribution is highly discussed. Models are often
high biased in the Northern Hemisphere leading to a NH / SH OH ratio greater than 1 (Naik et al., 2013). A methane
inversion using a NH / SH OH ratio
higher than 1 infers higher methane emissions in the Northern Hemisphere and
lower in the tropics and in the Southern Hemisphere (Patra et al., 2014).
However, there is recent evidence for parity in interhemispheric OH
concentrations (Patra et al., 2014), which needs to be confirmed by other
observational and modelderived estimates.
OH concentrations and their changes can be sensitive to climate variability
(e.g. Pinatubo eruption, Dlugokencky et al., 1996), to biomass burning
(Voulgarakis et al., 2015) and to anthropogenic activities. For instance, the
recent increase of the oxidizing capacity of the troposphere in South and East
Asia, associated with increasing NOxemissions and decreasing CO emissions (Mijling
et al., 2013; Yin et al., 2015), possibly enhances CH4 consumption
and therefore limits the atmospheric impact of increasing emissions (Dalsøren
et al., 2009). Despite such large regional changes, the global mean OH
concentration was suggested to have changed only slightly over the past 150
years (Naik et al., 2013). This is due to the concurrent increases of positive
influences on OH (water vapour, tropospheric ozone, nitrogen oxides (NOx ) emissions, and UV radiation due to
decreasing stratospheric ozone) and of OH sinks (methane burden, carbon
monoxide and non-methane volatile organic compound emissions and burden).
However the sign and integrated magnitude (from 1850 to 2000) of OH changes is
uncertain, varying from −13 to >+15 % among the ACCMIP models (mean of
−1 %, Naik et al., 2013). Dentener et al. (2003) found a positive trend
in global OH concentrations of 0.24 ± 0.06 % yr-1
between 1979 and 1993, mostly explained by changes in the tropical tropospheric
water vapour content. Accurate methyl chloroform atmospheric observations
together with estimates of its emissions (Montzka and Fraser, 2003) allow an
estimate of OH concentrations and changes in the troposphere from the 1980s.
Montzka et al. (2011) inferred small interannual OH variability and trends
(typical OH changes from year to year of less than 3 %) and attributed
previously estimated large year-to-year OH variations before 1998 (e.g.
Bousquet et al., 2005; Prinn et al., 2001) to overly large sensitivity of OH
concentrations inferred from methyl chloroform measurements to uncertainties in
the latter’s emissions. However, Prinn et al. (2005) also showed lower
post-1998 OH variability that they attributed to the lack of strong post-1998
El Niño events. For the ACCMIP models providing continuous simulations over the
past decades, OH interannual variability ranged from 0.4 to 0.9 %, consistent
but lower than the value deduced from methyl chloroform measurements. However
these runs take into account meteorology variability but not emission interannual
variability (e.g. from biomass burning) and thus are expected to simulate lower
OH interannual variability than in reality. As methyl chloroform has reached
very low concentrations in the atmosphere, in compliance with the application
of the Montreal Protocol and its amendments, a replacement compound is needed
to estimate global OH concentrations. Several hydrochlorofluorocarbons and
hydrofluorocarbons have been tested (Miller et al., 1998; Montzka et al., 2011;
Huang and Prinn, 2002) to infer OH but do not yet provide equivalent results to
methyl chloroform.
We report here a climatological range of 454– 617 Tg CH4 yr-1
as in Kirschke et al. (2013) for the total tropospheric loss of methane by OH
oxidation in the 2000s.
|
|
| 3.3.2.0 | Stratospheric loss |
| 3.3.2.0 | Stratospheric loss | | | | Approximately 60 Tg CH4 yr−1 enters the stratosphere by
cross-tropopause mixing and the Hadley circulation (Reeburgh, 2007).
Stratospheric CH4 distribution is highly correlated to the changes
in the Brewer–Dobson circulation (Holton, 1986) and may impact Arctic air
through subsidence of isotopically heavy air depending on the polar vortex
location (Röckmann et al., 2011). In the stratosphere, currently approximately
51 [16–84] Tg CH4 yr-1 (i.e. about 10 [3–16] % of the
total chemical loss in the atmosphere) is lost through reactions with excited
atomic oxygen O(1D), atomic chlorine (Cl), atomic fluorine (F) and
OH (Voulgarakis et al., 2013; Williams et al., 2012). The fraction of the
stratospheric loss due to the different oxidants is uncertain, possibly within
20–35 % due to halons, about 25 % due to O(1D), the rest being due
to stratospheric OH (Neef et al., 2010). The oxidation of methane in the
stratosphere produces significant amounts of water vapour, which has a positive
radiative forcing, and stimulates the production of OH through its reaction
with atomic oxygen (Forster et al., 2007). Stratospheric methane thus
contributes significantly to the observed variability and trend in
stratospheric water vapour (Hegglin et al., 2014). Uncertainties in the chemical
loss of stratospheric methane are large, due to uncertain interannual
variability in stratospheric transport as well as through its chemical
interactions with stratospheric ozone (Portmann et al., 2012).
We report here a climatological range of 16– 84 Tg CH4 yr−1 as in
Kirschke et al. (2013).
3.3.3
Tropospheric
reaction with Cl
Halogen
atoms can also contribute to the oxidation of methane in the troposphere. Allan
et al. (2005) measured mixing ratios of methane and δ13C–CH4 at two stations in the
Southern Hemisphere from 1991 to 2003 and found that the apparent kinetic
isotope effect of the atmospheric methane sink was significantly larger than
that explained by OH alone. A seasonally varying sink due to atomic chlorine
(Cl) in the marine boundary layer of between 13 and 37 Tg CH4 yr-1
was proposed as the explaining mechanism (Allan et al., 2007). This sink was
estimated to occur mainly over coastal and marine regions, where NaCl from
evaporated droplets of seawater react with NO2 to eventually form Cl2,
which then UV dissociates to Cl. However significant production of nitryl
chloride (ClNO2) at continental sites has been recently reported
(Riedel et al., 2014) and suggests the broader presence of Cl, which in turn
would expand the significance of the Cl sink in the troposphere. More work is
needed on this potential re-evaluation of the Cl impact on the methane budget.
We report here a climatological range of 13– 37
Tg CH4 yr-1 as in Kirschke et al. (2013).
|
|
| 3.3.4.0 | Soil uptake |
| 3.3.4.0 | Soil uptake | | | | Unsaturated oxic soils are sinks of atmospheric methane due to the presence
of methanotrophic bacteria, which consume methane as a source of energy.
Wetlands with temporally variable saturation can also act as methane sinks.
Dutaur and Verchot (2007) conducted a comprehensive meta-analysis of field
measurements of CH4 uptake spanning a variety of ecosystems. They
reported a range of 36 ± 23 Tg CH4 yr-1
but also showed that stratifying the results by climatic zone, ecosystem and
soil type led to a narrower range (and lower mean estimate) of 22 ±
12 Tg CH4 yr-1. A modelling study by Ridgwell et
al. (1999) simulated the sink to be 20– 51 Tg CH4 yr-1.
Curry (2007) used a process-based methane consumption scheme coupled to a land
surface model (and calibrated to field measurements) to obtain a global
estimate of 28 Tg CH4 yr-1, with a range of 9–47 Tg CH4
yr-1, which is the result reported in Kirschke et al. (2013). Tian
et al. (2016) further updated the CH4 uptake from soil, with the
estimate of 30 ± 19 Tg CH4 yr-1.
In that model, CH4 uptake was determined by the diffusion rate of
methane and oxygen through the uppermost soil layer, which was in turn
dependent upon the soil characteristics (e.g. texture, bulk density) and water
content (Curry, 2007). Riley et al. (2011) used another process-based model and
estimated a global atmospheric CH4 sink of 31 Tg CH4 yr-1.
The methane consumption rate was also dependent on the available soil water,
soil temperature and nutrient availability. Although not addressed in that model,
it should be noted that if the soil water content increases enough to inhibit
the diffusion of oxygen, the soil could become a methane source (Lohila et al.,
2016). This transition can be rapid, thus creating areas that can be either a
source or a sink of methane depending on the season.
Following Curry (2007), and consistent with Tian et al. (2015), we report
here a climatological range of 9–47 Tg CH4 yr-1 as in
Kirschke et al. (2013).
|
|
| 3.3.5.0 | CH4 lifetime |
| 3.3.5.0 | CH4 lifetime | | | | The global atmospheric lifetime is defined for a gas in steady state as the
global atmospheric burden (Tg) of this gas divided by its global total sink (Tg
yr-1) (IPCC, 2001). In a case of a gas whose local lifetime is
constant in space and time, the atmospheric lifetime equals the decay time (efolding time) of
a perturbation. As methane is not in a steady state, we need to fit with a
function that approaches steady state when calculating methane lifetime using
atmospheric measurements (Sect. 4.1.1). Global models provide an estimate of
the loss of the gas due to individual sinks, which can then be used to derive
lifetime due to a specific sink. For example, methane’s tropospheric lifetime
is determined as global atmospheric methane burden divided by the loss from OH
oxidation in the troposphere, sometimes called “chemical lifetime”, while its
total lifetime corresponds to the global burden divided by the total loss
including tropospheric loss from OH oxidation, stratospheric chemistry and soil
uptake. Recent multimodel estimate of the tropospheric methane lifetime is of
9.3 years (range 7.1–10.6; Voulgarakis et al., 2013; Kirschke et al., 2013) and
that of the total methane lifetime is 8.2 ± 0.8
years (for year 2000, range 6.4–9.2, Voulgarakis et al., 2013). The model
results for total methane lifetime are consistent with, though smaller than,
the value reported in Table 6.8 of the IPCC AR5 of 9.1 ± 0.9
years (which was the observationally constrained estimate of Prather et al.,
2012) most commonly used in the literature (Ciais et al., 2013) and the
steady-state calculation from atmospheric observations (9.3 years, Sect.
4.1.1).
|
|
| 4.0.0.0 | Atmospheric observations and top-down inversions |
| 4.0.0.0 | Atmospheric observations and top-down inversions | | | | |
|
| 4.1.0.0 | Atmospheric observations |
| 4.1.0.0 | Atmospheric observations | | | | The first systematic atmospheric CH4 observations began in 1978
(Blake et al., 1982) with infrequent measurements from discrete air samples
collected in the Pacific at a range of latitudes from 67◦ N to 53◦
S. Because most of these air samples were from well-mixed oceanic air masses
and the measurement technique was precise and accurate, they were sufficient to
establish an increasing trend and the first indication of the latitudinal
gradient of methane. Spatial and temporal coverage was greatly improved soon
after (Blake and Rowland, 1986) with the addition of the NOAA flask network
(Steele et al., 1987; Fig. 1), and of AGAGE (Cunnold et al., 2002), CSIRO
(Francey et al., 1999), and other networks (e.g. ICOS network in Europe, https://www.icos-ri.eu/). The
combined datasets provide the longest time series of globally averaged CH4
abundance. Since the early 2000s, remotely sensed retrievals of CH4
have provided CH4 atmospheric columnaveraged mole fractions
(Buchwitz et al., 2005a; Frankenberg et al., 2005; Butz et al., 2011;
Crevoisier et al., 2009; Wunch et al., 2011). Fourier transform infrared (FTIR)
measurements at fixed locations also provide methane column observations (Wunch
et al., 2011).
|
|
| 4.1.1.0 | In situ CH4 observations and atmospheric growth rate at the surface |
| 4.1.1.0 | In situ CH4 observations and atmospheric growth rate at the surface | | | | Four observational networks provide globally averaged CH4 mole
fractions at the Earth’s surface: the Earth System Research Laboratory from US
National Oceanic and Atmospheric Administration (NOAA/ESRL, Dlugokencky et al.,
1994), the Advanced Global Atmospheric Gases Experiment (AGAGE, Prinn et al.,
2000; Cunnold et al., 2002; Rigby et al., 2008), the Commonwealth Scientific
and Industrial Research Organisation (CSIRO, Francey et al., 1999) and the
University of California Irvine (UCI, Simpson et al., 2012). The data are
archived at the World Data Centre for Greenhouse Gases (WDCGG) of the WMO
Global Atmospheric Watch (WMO-GAW) programme, including measurements from other
sites that are not operated as part of the four networks.
The networks differ in their sampling strategies, including the frequency of
observations, spatial distribution, and methods of calculating globally
averaged CH4 mole fractions. Details are given in the Supplement of
Kirschke et al. (2013). For the global average values of CH4
concentrations presented here, all measurements are made using gas
chromatography with flame ionization detection (GC/FID), although
chromatographic schemes vary among the labs. Because GC/FID is a relative measurement
method, the instrument response must be calibrated against standards. NOAA
maintains the WMO CH4 mole fraction scale X2004A; NOAA and CSIRO
global means are on this scale. AGAGE uses an independent standard scale (Aoki
et al., 1992), but direct comparisons of standards and indirect comparisons of
atmospheric measurements show that differences are below 5 ppb (WMO RoundRobin
programme). UCI uses another independent scale that was established in 1978 and
is traceable to NIST (Simpson et al., 2012) but has not been included in
standard exchanges with other networks so differences with the other networks
cannot be quantitatively defined. Additional experimental details are presented
in the Supplement from Kirschke et al. (2013) and references therein.
In Fig. 1, (a) globally averaged CH4 and (b) its growth rate
(derivative of the deseasonalized trend curve) through 2012 are plotted for a
combination of the four measurement programmes using a procedure of signal
decomposition described in Thoning et al. (1989). We define the annual increase
GATMas the increase in the growth rate from 1
January in one year to 1 January in the next year. Agreement among the four
networks is good for the global growth rate, especially since ∼
1990. The long-term behaviour of globally averaged atmospheric CH4
shows a decreasing but positive growth rate (defined as the derivative of the
deseasonalized mixing ratio) from the early 1980s through 1998, a
near-stabilization of CH4 concentrations from 1999 to 2006, and a
renewed period with positive but stable growth rates since 2007. When a
constant atmospheric lifetime is assumed, the decreasing growth rate from 1983
through 2006 implies that atmospheric CH4 was approaching steady
state, with no trend in emissions. The NOAA global mean CH4
concentration was fitted with a function that describes the approach to a
first-order steady state (SSindex): [CH4](t ) = [CH4]SS
− ([CH4]SS −
[CH4]0)e−t/τ;
solving for the lifetime, τ , gives 9.3 years, which is very close to current
literature values (e.g. Prather et al., 2012).
On decadal timescales, the annual increase is on average 2.1 ±
0.3 ppb yr-1 for 2000–2009, 3.5 ± 0.2
ppb yr-1 for 2003–2012 and 5.0 ± 1.0
ppb yr-1 for the year 2012. The two decadal values hide a jump in
the growth rate after 2006. Indeed, from 1999 to 2006, the annual increase of
atmospheric CH4 was remarkably small at 0.6 ± 0.1
ppb yr-1. In the last 8 years, the atmospheric growth rate has
recovered to a level similar to that of the mid-1990s (∼ 5
ppb yr-1), before the stabilization period of 1999–2006, as stated
in Kirschke et al. (2013).
|
|
| 4.1.2.0 | Satellite data of column-averaged CH4 |
| 4.1.2.0 | Satellite data of column-averaged CH4 | | | | In the 2000s, two space-borne instruments sensitive to atmospheric methane
were put in orbit and have provided atmospheric methane column-averaged dry air
mole fraction (XCH4), using either shortwave infrared spectrometry
(SWIR) or thermal infrared spectrometry (TIR).
Between 2003 and 2012, the Scanning Imaging Absorption spectrometer for
Atmospheric CartograpHY (SCIAMACHY) was operated on board the ESA ENVIronmental
SATellite (ENVISAT), providing nearly 10 years of XCH4 sensitive to
the atmospheric boundary layer (Burrows et al., 1995; Buchwitz et al., 2006;
Dils et al., 2006; Frankenberg et al., 2011). These satellite retrievals were
the first to be used for global and regional inverse modelling of methane
fluxes (Meirink et al., 2008a; Bergamaschi et al., 2007, 2009). The relatively
long time record allowed the analysis of the interannual methane variability
(Bergamaschi et al., 2013). However, the use of SCIAMACHY necessitates
important bias correction, especially after 2005 (up to 40 ppb from south to
north) (Bergamaschi et al., 2009; Houweling et al., 2014; Alexe et al., 2015).
In January 2009, the JAXA satellite Greenhouse Gases Observing SATellite
(GOSAT) was launched containing the TANSO-FTS instrument, which observes in the
shortwave infrared (SWIR). Different retrievals of methane based on
TANSO-FTS/GOSAT products are made available to the community (Yoshida et al., 2013;
Schepers et al., 2012; Parker et al., 2011) based on two retrieval approaches:
proxy and full physics. The proxy method retrieves the ratio of methane column
(XCH4) and carbon dioxide column (XCO2), from which XCH4
is derived after multiplication with transport model-derived XCO2 (Chevallier
et al., 2010; Peters et al., 2007; Frankenberg et al., 2006). It intends mostly
to remove biases due to light scattering on clouds and aerosols and is highly
efficient owing to the small spectral distance between CO2 and CH4
sunlight absorption bands (1.65 µm for CH4 and 1.60 µm for CO2).
Because of this, scattering-induced errors are similar for XCO2 and XCH4
and cancel out in the ratio. The second approach is the full-physics algorithm,
which retrieves the aerosol properties (amount, size and height) along with CO2 and CH4
columns (e.g. Butz et al., 2011). Although GOSAT retrievals still show
significant unexplained biases (possibly also linked to atmospheric transport
modelling; Locatelli et al., 2015) and limited sampling in cloud-covered
regions and in the high-latitude winter, it represents an important improvement
compared to SCIAMACHY both for random and systematic observation errors (see
Table S2 of Buchwitz et al., 2016).
Atmospheric inversions based on SCIAMACHY or GOSAT CH4 retrievals
have been carried out by different research groups (Monteil et al., 2013;
Cressot et al., 2014; Alexe et al., 2015; Bergamaschi et al., 2013; Locatelli
et al., 2015). For GOSAT, differences between the use of proxy and full-physics
retrievals have been investigated. In addition, joint CO2–CH4
inversions have been conducted to investigate the use of GOSAT retrieved ratios
avoiding a modelderived hard constraint on XCO2 (Pandey et al.,
2015, 2016; Fraser et al., 2013). Results from some of these studies are
reported in Sect. 5 of this paper.
|
|
| 4.1.3.0 | Methane isotope observations |
| 4.1.3.0 | Methane isotope observations | | | | The processes emitting methane discriminate differently its isotopologues
(isotopes). The two main stable isotopes of CH4 are 13CH4
and CH3D,and there is also the radioactive carbon
isotope 14C–CH4. Isotopic signatures are
conventionally given by the deviation of the sample mole ratio (for example, R=13CH4/12CH4or
CH3D/ CH4)
relative to a given standard (Rstd)relative to a reference ratio, given in per
mil as in Eq. (3).
δ13CH4 or
δD (CH4) >= − 1 ×
1000 (3)Rstd
For the 13CH4 isotope, the conventional
reference standard is known as Vienna Pee Dee Belemnite (VPDB), with Rpdb = 0.0112372. The same definition applies to CH3D,with the Vienna Standard Mean Ocean Water
(VSMOW) RSMOW= 0.00015575.
The isotopic composition of atmospheric methane is measured at a subset of
surface stations (Quay et al., 1991, 1999; Lowe et al., 1994; Miller et al.,
2002; Morimoto et al., 2006; Tyler et al., 2007). The mean atmospheric values
are about −47 ‰ for δ13CH4
and −86/−96 ‰ for δD(CH4).
Isotopic measurements are made mainly on flask air samples analysed with
gaschromatograph isotope ratio spectrometry for which an accuracy of 0.05 ‰ for
δ13CH4
and 1.5 ‰ for δD(CH4)
can be achieved (Rice et al., 2001; Miller et al., 2002). These isotopic
measurements based on air flask sampling have relatively low spatial and
temporal resolutions. Laser-based absorption spectrometers and isotope ratio
mass spectrometry techniques have recently been developed to increase sampling
frequency and allow in situ operation (McManus et al., 2010; Santoni et al.,
2012).
Measurements of δ13CH4
can help to partition the different methanogenic processes of methane: biogenic
(−70 to −55 ‰), thermogenic (−55 to −25 ‰) or pyrogenic
(−25 to −15 ‰) sources (Quay et al., 1991; Miller et al., 2002;
Fisher et al., 2011) or even the methanogenic pathway (McCalley et al., 2014). δD(CH4) provides
valuable information on the oxidation by the OH radicals (Röckmann et al.,
2011) due to a fractionation of about 300 ‰. Emissions also show substantial
differences in δD(CH4)
isotopic signatures: −200 ‰ for biomass burning sources vs. −360 to
−250 ‰ for biogenic sources (Melton et al., 2012; Quay et al., 1999). 14C–
CH4 measurements (Quay et al., 1991, 1999; Lowe et al., 1988) may
also help to partition for fossil fuel contribution (radiocarbon-free source).
For example, Lassey et al. (2007a) used more than 200 measurements of
radioactive 14C–CH4 (with a balanced weight between
Northern and Southern hemispheres) to further constrain the fossil fuel
contribution to the global methane source emission to 30 ±
2 % for the period 1986–2000.
Integrating isotopic information is important to improve our understanding
of the methane budget. Some studies have simulated such isotopic observations
(Neef et al., 2010; Monteil et al., 2011) or used them as additional
constraints to inverse systems (Mikaloff Fletcher et al., 2004; Hein et al.,
1997; Bousquet et al., 2006; Neef et al., 2010; Thompson et al., 2015). Using
pseudo-observations, Rigby et al. (2012) found that quantum-cascade-laser-based
isotopic observations would reduce the uncertainty in four major source
categories by about 10 % at the global scale (microbial, biomass burning,
landfill and fossil fuel) and by up to 50 % at the local scale. Although all
source types cannot be separated using 13C, D and 14C
isotopes, such data bring valuable information to constrain groups of sources
in atmospheric inversions, if the isotopic signatures of the various sources
can be precisely assessed (Bousquet et al., 2006, Supplement).
|
|
| 4.1.4.0 | Other atmospheric observations |
| 4.1.4.0 | Other atmospheric observations | | | | Other types of methane measurements are available, which are not commonly
used to infer fluxes from inverse modelling (yet) but are used to verify its
performance (see e.g. Bergamaschi et al., 2013). Aircraft or balloon-borne in
situ measurements can deliver vertical profiles with high verti cal resolution.
Such observations can also be used to test remote-sensing measurement from
space or from the surface and bring them on the same scale as the in situ
surface mea surements. Aircraft measurements have been undertaken in various
regions either during campaigns (Wofsy, 2011; Beck et al., 2012; Chang et al.,
2014; Paris et al., 2010) or in a recurrent mode using small aircrafts in the
planetary bound ary layer (Sweeney et al., 2015; Umezawa et al., 2014; Gatti et
al., 2014) and commercial aircrafts (Schuck et al., 2012; Brenninkmeijer et
al., 2007; Umezawa et al., 2012, 2014; Machida et al., 2008). Balloons can
carry in situ instruments (e.g. Joly et al., 2008; using tunable laser diode
spectrome try) or air samplers (e.g. air cores, Karion et al., 2010) up to 30
km height. New technologies have also developed systems based on cavity
ring-down spectroscopy (CRDS), opening a large ensemble of new activities to
estimate methane emissions such as drone measurements (light version of CRDS),
as land-based vehicles for real-time, mobile monitoring over oil and gas
facilities, as well as ponds, landfills, livestock, etc.
In October 2006, the Infrared Atmospheric Sounding In terferometer (IASI) on
board the European MetOp-A satel lite began to operate. Measuring the thermal
radiation from Earth and the atmosphere in the TIR, it provides mid-to upper
troposphere columns of methane (representative of the 5–15 km layer) over the
tropics using an infrared sounding interferometer (Crevoisier et al., 2009).
Despite its sensitivity being limited to the mid-to-upper troposphere, its use
in flux inversions has shown consistent results in the tropics with surface and
other satellite-based inversions (Cressot et al., 2014).
The Total Carbon Column Observing Network (TCCON) uses ground-based Fourier
transform spectrometers to measure atmospheric column abundances of CO2,,
CO, CH4, N2O
and other molecules that absorb sunlight in the near infrared spectral region
(Wunch et al., 2011). As TCCON measurements make use of sunlight, they can be
performed throughout the day during clear-sky conditions, with the sun
typically 10◦ above the horizon. The TCCON network has been established
as a reference for the validation of column retrievals, like those from
SCIAMACHY and GOSAT. TC CON data can be obtained from the TCCON Data Archive,
hosted by the Carbon Dioxide Information Analysis Center (CDIAC, http://cdiac.ornl.gov/).
|
|
| 4.2.0.0 | Top-down inversions |
| 4.2.0.0 | Top-down inversions | | | | |
|
| 4.2.1.0 | Principle of inversions |
| 4.2.1.0 | Principle of inversions | | | | An atmospheric inversion for methane fluxes (sources and sinks) optimally
combines atmospheric observations of methane and associated uncertainties, a
prior knowledge of the fluxes including their uncertainties, and a chemistry
transport model to relate fluxes to concentrations (Rodgers, 2000). In this
sense, top-down inversions integrate all the components of the methane cycle
described previously in this paper. The observations can be surface or
upper-air in situ observations, as well as satellite and surface retrievals.
Prior emissions generally come from bottom-up approaches such as process-based
models or data-driven extrapolations (nat ural sources) and inventories
(anthropogenic sources). The chemistry transport model can be Eulerian or
Lagrangian, and global or regional, depending on the scale of the flux to be
optimized. Atmospheric inversions generally rely on the Bayes’ theorem, which
leads to the minimization of a cost function as Eq. (4): (See original
document)
where y is
a vector containing the atmospheric observations, x is a state vector containing the
methane emissions and other appropriate variables (like OH concentrations or CH4
concentrations at the start of the assimilation window) to be estimated, xbis the prior state of x, and H is
the observation operator, here the combination of an atmospheric transport and
chemistry model and an interpolation procedure sam pling the model at the
measurement coordinates. R is the er ror covariance
matrix of the observations and Pb is the error covariance matrix associated with xb. The errors on the mod elling of atmospheric transport and
chemistry are included in the R matrix (Tarantola, 1987).
The minimization of a lin earized version of J leads to the optimized state
vector xa(Eq. 5): (See original document) where Pa is
given by Eq. (6) and represents the error covari ance matrix associated with xa, and H contains the sensitiv ities
of any observation to any component of state vector x (linearized version of the
observation operator H (x)).(See original document)
Unfortunately, the size of the inverse problem usually does not allow
computing Pa , which is therefore approximated
using the leading eigenvectors of the Hessian of J (Chevallier et al., 2005) or from stochastic ensembles
(Chevallier et al., 2007). Therefore, the optimized fluxes xaare
obtained using classical minimization algorithms (Chevallier et al., 2005;
Meirink et al., 2008b). Alternatively, Chen and Prinn (2006) computed monthly
emissions by applying a recursive Kalman filter in which Pa is
computed explicitly for each month. Emissions are generally derived at weekly
to monthly timescales, and for spatial resolutions ranging from model grid
resolution to large aggregated regions. Spatiotemporal aggregation of state
vector elements reduces the size of the inverse problem and allows the
computation of Pa . However, such
aggregation can also generate aggregation errors inducing possible biases in
the inferred emissions and sinks (Kaminski et al., 2001). The estimated xacan represent either the net methane flux
in a given region or contributions from specific source categories. Atmospheric
inversions use bottom-up models and inventories as prior estimates of the
emissions and sinks in their setup, which make bottom-up and top-down
approaches generally not independent.
|
|
| 4.2.2.0 | Reported inversions |
| 4.2.2.0 | Reported inversions | | | | A group of eight atmospheric inversion systems using global Eulerian
transport models were used in this synthesis. Each inversion system provides
from 1 to 10 inversions, including sensitivity tests varying the assimilated
observations (surface or satellite) or the inversion setup. This represents a
total of 30 inversion runs with different time coverage: generally 2000–2012
for surface-based observations, 2003–2012 for SCIAMACHY-based inversions and 2009–2012
for GOSATbased inversions (Table 3). When multiple sensitivity tests were
performed we use the mean of this ensemble not to overweight one particular
inverse model. Bias correction procedures have been developed to assimilate
SCIAMACHY (Bergamaschi et al., 2009, 2013; Houweling et al., 2014) and GOSAT
data (Cressot et al., 2014; Houweling et al., 2014; Locatelli et al., 2015;
Alexe et al., 2015). These procedures can lead to corrections from several
parts per billion and up to several tens of parts per billion (Bergamaschi et
al., 2009; Locatelli et al., 2015). Although partly due to transport model
errors, the large corrections applied to satellite total column CH4
data question the comparably low systematic errors reported in satellite
validation studies using TCCON (Dils et al., 2014; CCI-Report, 2016). It should
also be noticed that some satellite-based inversions are in fact combined
satellite and surface inversions as they use either instantaneous in situ data
simultaneously (Bergamaschi et al., 2013; Alexe et al., 2015) or annual mean
surface observations to correct satellite bias (Locatelli et al., 2015).
Nevertheless, these inversions are still referred to as satellite-based
inversions.
General characteristics of the inversion systems are provided in Table 3.
Further detail can be found in the referenced papers. Each group was asked to
provide gridded flux estimates for the period 2000–2012, using either surface
or satellite data, but no additional constraints were imposed so that each group
could use their preferred inversion setup. This approach is appropriate for our
purpose of flux assessment but not necessarily for model intercomparison. We
did not require posterior uncertainty from the different participating groups,
which may be done for the next release of the budget. Indeed chemistry
transport models have some limitations that impact on the inferred methane
budget, such as discrepancies in interhemispheric transport, stratospheric
methane profiles and OH distribution. We consider here an ensemble of
inversions gathering a large range of chemistry transport models, through their
differences in vertical and horizontal resolutions, meteorological forcings,
advection and convection schemes and boundary layer mixing; we assume that this
model range is sufficient to cover the range of transport model errors in the
estimate of methane fluxes. Each group provided gridded monthly maps of
emissions for both their prior and posterior total and for sources per category
(see the categories Sect. 2.3). Results are reported in Sect. 5. Atmospheric
sinks were not analysed for this budget, which still relies on Kirschke et al.
(2013) for bottom-up budget and on a global mass balance for top-down budget
(difference between the global source and the observed atmospheric increase).
The last year of reported inversion results is 2012, which represents a
4-year lag with the present. Satellite observations are linked to operational
data chains and are generally available within days to weeks after the
recording of the spectra. Surface observations can lag from months to years
because of the time for flask analyses and data checks in (mostly)
non-operational chains. With operational networks such as ICOS in Europe, these
lags will be reduced in the future. In addition, the final 6 months of
inversions are generally ignored (spun down) because the estimated fluxes are
not constrained by as many observations as the previous months. Finally, the
long inversion runs and analyses can take up to months to be performed. For the
next global methane budget the objective is to represent more recent years by
reducing the analysis time and shortening the in situ atmospheric observation
release.
|
|
| 5.0.0.0 | Methane budget: top-down and bottom-up comparison |
| 5.0.0.0 | Methane budget: top-down and bottom-up comparison | | | | |
|
| 5.1.0.0 | Global methane budget |
| 5.1.0.0 | Global methane budget | | | | |
|
| 5.1.1.0 | Global budget of total methane emissions |
| 5.1.1.0 | Global budget of total methane emissions | | | | |
|
| 5.1.1.1 | Top-down estimates |
| 5.1.1.1 | Top-down estimates | | | | At the global scale, the total emissions inferred by the ensemble of 30
inversions are 558 Tg CH4 yr-1 [540–570] for the
2003–2012 decade (Table 4), with a higher value of 568 Tg CH4 yr-1
[542–582] for 2012. Global emissions for 2000–2009 (552 Tg CH4 yr-1)
are consistent with Kirschke et al. (2013), and the range of uncertainties for
global emissions (535–566) is in line as well with that of Kirschke et al.
(2013) (526–569), although 8 out of the 30 inversions presented here (∼
25 %) are different. The latitudinal breakdown of emissions inferred from
atmospheric inversions reveals a dominance of tropical emissions at 359 Tg CH4
yr-1 [339– 386], representing 64 % of the global total. Thirty-two
per cent of the emissions are from the midlatitudes and 4 % from high latitudes
(above 60◦ N).
|
|
| 5.1.1.2 | Bottom-up estimates |
| 5.1.1.2 | Bottom-up estimates | | | | The picture given by the bottom-up approaches is quite different with global
emissions of 736 Tg CH4 yr-1 [596–884] for 2003–2012
(Table 2). This estimate is much larger than top-down estimates. The bottom-up
estimate is given by the sum of individual anthropogenic and natural processes,
with no constraint on the total. As noticed in Kirschke et al. (2013), such a
large global emissions rate is not consistent with atmospheric constraints
brought by OH optimization and is very likely overestimated. This
overestimation likely results from errors in the estimation of natural sources
and sinks: extrapolation or double counting of some natural sources (e.g.
wetlands, inland waters), or estimation of atmospheric sink terms. The
anthropogenic sources are much more consistent between bottom-up and top-down
approaches (Sect. 5.1.2).
|
|
| 5.1.2.0 | Global methane emissions per source category |
| 5.1.2.0 | Global methane emissions per source category | | | | The global methane budget for five source categories (see Sect. 2.3) for
2003–2012 is presented in Fig. 5 and Table 2. Top-down estimates attribute
about 60 % of the total emissions to anthropogenic activities (range of 50–70
%) and 40 % to natural emissions. As natural emissions from bottom-up models
are much larger, the anthropogenic vs. natural emission ratio is more balanced
for bottom-up (∼ 50 % each). A predominant role of anthropogenic
sources of methane emissions is strongly supported by the ice core and
atmospheric methane records. The data indicate that atmospheric methane varied
around 700 ppb during the last millennium before increasing by a factor of 2.6
to ∼
1800 ppb. Accounting for the decrease in mean lifetime over the industrial
period, Prather et al. (2012) estimate from these data a total source of 554 ±
56 Tg CH4 in 2010 of which about 64 % (352 ±
45 Tg CH4) are of anthropogenic origin, very consistent
estimates with our synthesis.
|
|
| 5.1.2.1 | Wetlands |
| 5.1.2.1 | Wetlands | | | | For 2003–2012, the top-down and bottom-up derived estimates of respectively
167 Tg CH4 yr-1 (range 127–202) and 185 Tg CH4
yr-1 (range 153–227) are statistically consistent. Mean wetland
emissions for the 2000–2009 period appear similar, albeit slightly smaller than
found in Kirschke et al. (2013): 166 Tg CH4 yr−1 in this
study vs. 175 Tg CH4 yr-1 in Kirschke et al. (2013) for
top-down (−4 %) and 183 Tg CH4 yr-1 in this study
vs. 217 Tg yr-1 in Kirschke et al. (2013) for bottom-up (−15
%). Note that more inversions (top-down) and more wetland models
(bottom-up) were used in this study. Inversions have difficulty in
separating wetlands from other sources so that uncertainties on top-down
wetland emissions remain large. In this study, all bottom-up models were forced
with the same wetland extent and climate forcings (Poulter et al., 2016), with
the result that the amplitude of the range of emissions of 151–222 for
2000–2009 has narrowed by a third compared to the previous estimates from
Melton et al. (2013) (141–264) and from Kirschke et al. (2013) (177–284). This
suggests that differences in wetland extent explain about a third (30–40 %) of
the former range of the emission estimates of global natural wetlands. The
remaining range is due to differences in model structures and parameters. It is
also worth noting that bottom-up and top-down estimates differ less in this
study (∼ 17 Tg yr-1 for the mean) than in
Kirschke et al. (2013) (∼ 30 Tg yr-1),
although results from many more models are reported here. For top-down
inversions, natural wetlands represent 30 % on average of the total methane
emissions but only 25 % for bottom-up models (because of higher total emissions
inferred by bottom-up models).
|
|
| 5.1.2.2 | Other natural emissions |
| 5.1.2.2 | Other natural emissions | | | | The discrepancy between top-down and bottom-up budgets is the largest for
the natural emission total, which is 384 Tg CH4 yr-1
[257–524] for bottom-up and only 231 Tg CH4 yr-1
[194–296] for top-down over the 2003– 2012 decade. Processes other than natural
wetlands (Fig. 5), namely freshwater systems, geological sources, termites, oceans,
wild animals, wildfires, and permafrost, explain this large discrepancy. For
the 2003–2012 decade, topdown inversions infer non-wetland natural emissions of
64 Tg CH4 yr-1 [21–132], whereas the sum of the
individual bottom-up emissions is 199 Tg CH4 yr-1
[104–297]. The two main contributors to this large bottom-up total are
freshwater (∼ 60 %) and geological emissions (∼
20 %), both of which have large uncertainties without spatially explicit
representation. Because of the discrepancy, this category represents 10 % of
total emissions for top-down inversions but 27 % for bottom-up approaches.
Improved area estimates of freshwater emissions would be beneficial. For
example, stream fluxes are difficult to assess because of the high-expected
spatial variability and very uncertain areas of headwater streams where
methane-rich groundwater may be rapidly degassed. There are also uncertainties
in the geographical distinction between wetlands, small lakes (e.g. thermokarst
lakes), and floodplains that will need more attention to avoid double counting.
In addition, major uncertainty is still associated with representation of
ebullition. The intrinsic nature of this large but very locally distributed
flux highlights the need for cost-efficient highresolution techniques for
resolving the spatio-temporal variations of these fluxes. In this context of
observational gaps in space and time, freshwater fluxes are considered
underestimated until measurement techniques designed to properly account for
ebullition become more common (Wik et al., 2016a). On the contrary, global
estimates for freshwater emissions rely on upscaling of uncertain emission
factors and emitting areas, with probable overlapping of wetland emissions
(Kirschke et al., 2013), which may also lead to an overestimate. More work is
needed, based on both observations and process modelling, to overcome these
uncertainties.
For geological emissions, relatively large uncertainties come from the
extrapolation of only a subset of direct measurements to estimate the global
fluxes. Moreover, marine seepage emissions are still widely debated (Berchet et
al., 2016), and particularly diffuse emissions from microseepage are highly
uncertain. However, summing up all fossilCH4-related sources
(including the anthropogenic emissions) leads to a total of 173 Tg CH4
yr−1 [149–209],
which is about 31 % [25–35 %] of global methane emissions. This result is
consistent with 14C atmospheric isotopic analyses inferring a 30 %
contribution of fossil-CH4 to global emissions (Lassey et al.,
2007b; Etiope et al., 2008). All nongeological and non-wetland land source
categories (wild animals, wildfires, termites, permafrost) have been evaluated
at a lower level than in Kirschke et al. (2013) and contribute only 23 Tg CH4
yr-1 [9–36] to global emissions. From a topdown point of view, the
sum of all natural sources is more robust than the partitioning between
wetlands and other natural sources. To reconcile top-down inversions and
bottom up estimates, the estimation and proper partition of methane emissions
from wetlands and freshwater systems should receive high priority.
|
|
| 5.1.2.3 | Anthropogenic emissions |
| 5.1.2.3 | Anthropogenic emissions | | | | Total anthropogenic emissions are found statistically consistent between
top-down (328 Tg CH4 yr-1, range 259–370) and bottom-up
approaches (352 Tg CH4 yr-1, range 340– 360), although
top-down average is about 7 % smaller than bottom-up average over 2003–2012.
The partition of anthropogenic emissions between agriculture and waste, fossil
fuel extraction and use, and biomass and biofuel burning also shows good
consistency between top-down and bottom-up approaches (Table 2 and Fig. 7). For
2003–2012, agriculture and waste contributed 188 Tg CH4 yr-1
[115–243] for top-down and 195 Tg CH4 yr-1 [178–206] for
bottom-up. Fossil fuel emissions contributed 105 Tg CH4 yr-1
[77–133] for top-down and 121 Tg CH4 yr-1 [114–133] for
bottomup. Biomass and biofuel burning contributed 34 Tg CH4 yr-1
[15–53] for top-down and 30 Tg CH4 yr-1 [27–35] for
bottom-up. Biofuel methane emissions rely on very few estimates at the moment
(Wuebbles and Hayhoe, 2002; GAINS model). Although biofuel is a small source
globally (∼ 12 Tg CH4 yr-1), more
estimates are needed to allow a proper uncertainty assessment. Overall for
top-down inversions the global fraction of total emissions for the different
source categories are 33 % for agriculture and waste, 20 % for fossil fuels,
and 6 % for biomass and biofuel burnings. With the exception of biofuel
emissions, the global uncertainty of anthropogenic emissions appears to be
smaller than that of natural sources but with asymmetric uncertainty
distribution (mean significantly different than median). In poorly observed
regions, top-down inversions rely on the prior estimates and bring little or no
additional information to constrain the (often) spatially overlapping emissions
(e.g. in India, China). Therefore, the relative agreement between top-down and
bottom-up may indicate the limited capability of the inversion to separate the
emissions and should therefore be treated with caution. Although the
uncertainty range of some emissions has been decreased in this study compared
to Kirschke et al. (2013) (e.g. oceans, termites, geological), there is no
uncertainty reduction in the regional budgets because of the larger range
reported for emissions from freshwater systems.
|
|
| 5.2.0.0 | Regional methane budget |
| 5.2.0.0 | Regional methane budget | | | | |
|
| 5.2.1.0 | Regional budget of total methane emissions |
| 5.2.1.0 | Regional budget of total methane emissions | | | | At regional scale, for the 2003–2012 decade (Table 4 and Fig. 6), total
methane emissions are dominated by Africa with 86 Tg CH4 yr-1
[73–108], tropical South America with a total of 84 Tg CH4 yr-1
[65–101], and South East Asia with 73 Tg CH4 yr-1
[55–84]. These three (mainly) tropical regions represent almost 50 % of methane
emissions worldwide. The other high-emitting source regions are China (58 Tg CH4
yr-1 [51-72]), central Eurasia and Japan (46 Tg CH4 yr-1
[38–54]), contiguous USA (41 Tg CH4 yr-1 [34–49]), Russia
(38 Tg CH4 yr-1 [31–44]), India (39 Tg CH4 yr-1
[37–46]) and Europe (28 Tg CH4 yr-1 [21–34]). The other
regions (boreal and central North America, temperate South America, Oceania,
oceans) contribute between 7 and 20 Tg CH4 yr-1. This
budget is consistent with Kirschke et al. (2013) within the large ranges around
the mean emissions, although larger emissions are found here for South America,
South East Asia, and Europe and lower emissions are found for Africa, North
America and China. The regions with the largest changes are usually the least
constrained by the surface networks.
The different inversions assimilated either satellite or ground-based
observations. It is of interest to determine whether these two types of data
provide consistent surface emissions. To do so, we computed global, hemispheric
and regional methane emissions using satellite-based inversions and
ground-based inversions separately for the 2010–2012 time period, which is the
longest time period for which results from both GOSAT satellite-based and
surface-based inversions were available. At the global scale, satellite-based
inversions infer significantly higher emissions (>+12 Tg CH4 yr-1,
p >= 0.04)
than ground-based in versions. At the regional scale, emissions varied between
the satellite-based and surface-based inversions, although the difference is
not statistically significant due to too few inversions and some outliers
making the ensemble not robust enough. Yet the largest differences
(satellite-based minus surface based inversions) are observed over the tropical
region: tropical South America >+11 Tg CH4 yr-1; southern
Africa >+6 Tg CH4 yr-1; India −6 Tg CH4
yr-1; and over China −7 Tg CH4 yr-1.
Satellite data provide more constraints on fluxes in tropical regions than
surface-based inversions, due to a much larger spatial coverage. It is
therefore not surprising that most differences between these two types of
observations are found in the tropical band. However, such differences could
also be due to the larger systematic errors of satellite data as compared to
surface networks (Dils et al., 2014). In this context, the way the stratosphere
is treated in the atmospheric models used to produce atmospheric methane
columns from remote-sensing measurements (e.g. GOSAT or TCCON) seems important
to further investigate (Locatelli et al., 2015; Monteil et al., 2011; Bergamaschi
et al., 2009). Recent papers have developed methodologies to extract
tropospheric partial column abundances from the TCCON data (Saad et al., 2014;
Wang et al., 2014). Such partitioning could help explain the discrepancies
between atmospheric models and satellite data.
|
|
| 5.2.2.0 | Regional methane emissions per source category |
| 5.2.2.0 | Regional methane emissions per source category | | | | The analysis of the regional methane budget per source category (Fig. 7) can
be performed both for bottom-up and top-down approaches but with limitations. A
complementary view of the methane budget is also available as an interactive
graphic produced using data visualization techniques
(http://lsce-datavisgroup.github.io/MethaneBudget/). Moving the mouse over
regions, processes or fluxes reveals their relative weights in the global
methane budget and provides the mean values and the minimum–maximum ranges of
their contributions (mean [min, max]). The total source estimates from the
bottom-up approaches are further classed into finer subcategories. This graphic
shows that there is good consistency between top-down and bottom-up approaches
in the partition of anthropogenic emissions between agriculture and waste,
fossil fuel extraction and use, and biomass and biofuel burning, and it also
highlights the disequilibrium between top-down (left) and bottom-up (right)
budgets, mainly due to natural sources. On the bottom-up side, some natural
emissions are not (yet) available at regional scale (oceans, geological, inland
waters). Therefore, the category “others” is not shown for bottom-up results in
Fig. 7 and is not regionally attributed in the interactive graphic. On the
top-down side, as already noted, the partition of emissions per source category
has to be considered with caution. Indeed, using only atmospheric methane observations
to constrain methane emissions makes this partition largely dependent on prior
emissions. However, differences in spatial patterns and seasonality of
emissions can still be constrained by atmospheric methane observations for
those inversions solving for different sources categories (see Sect. 2.3).
Wetland emissions largely dominate methane emissions in tropical South
America, boreal North America, southern Africa, temperate South America and
South East Asia, although agriculture and waste emissions are almost as
important for the last two regions. Agriculture and waste emissions dominate in
India, China, contiguous USA, central North America, Europe and northern
Africa. Fossil fuel emissions dominate in Russia and are close to agriculture
and waste emissions in the region called central Eurasia and Japan. In China,
fossil fuel emissions are on average close, albeit smaller, than agriculture
and waste emissions. Comparison between bottom-up and top-down approaches shows
good consistency, but one has to consider the generally large error bars,
especially for top-down inversions. The largest discrepancy occurs for wetland
emissions in boreal North America where bottom-up models infer larger emissions
(32 Tg CH4 yr-1) than top-down inversions (13 Tg CH4
yr-1). Indeed, one particular bottom-up model infers a 61 Tg CH4
yr-1 emission for this region, largely above estimates from other
models, which lie between 15 and 45 Tg CH4 yr-1. Top-down
models results are consistent with the climatology proposed by Kaplan (2002),
whereas bottom-up models are more in line, albeit larger, than the climatology
of Matthews and Fung (1987), who infer about 30 Tg CH4 yr-1
for boreal North America. Interestingly, the situation is different for Russia
where top-down and bottomup approaches show similar mean emissions from natural
wetlands (mostly boreal, ∼ 13–14 Tg CH4 yr-1),
consistently with Kaplan (2002) but not with Matthews and Fung (1987), who
infer almost 50 Tg CH4 yr−1 for Russia.
Wetland emissions from Russia appear very uncertain, as also found by Bohn
et al. (2015) for western Siberia. Wetland emissions from tropical South
America are found more consistent in this work than in Kirschke et al. (2013),
where topdown inversions showed 2 times less emission than bottomup models. The
larger number of bottom-up models (11 against 3) and top-down inversions (30
against 8) are plausible causes explaining the improved agreement in this
tropical region, poorly constrained by the surface networks (Pison et al.,
2013).
Anthropogenic emissions remain close between top-down and bottom-up
approaches for most regions, again with the possibility that part of this
agreement is due to the lack of information brought by atmospheric observations
to top-down inversions for some regions. One noticeable exception is the lower
emissions for China as compared to the prior, visible also in Fig. 6. A priori
anthropogenic emissions for China are mostly provided by the EDGARv4.2
inventory. Starting from prior emissions of 67 Tg CH4 yr-1
[58–77], the mean of the atmospheric derived estimates for China is 58 Tg CH4
yr-1 [51–72], corresponding to a −14 % difference of the
Chinese emissions. A t test
performed for the available estimates suggests that the mean posterior total
emission for China is different from the prior emission at the 95 % confidence
level. Several atmospheric studies have already suggested a possible
overestimation of methane emissions from coal in China in the EDGARv4.2
inventory (Bergamaschi et al., 2013; Kirschke et al., 2013; Tohjima et al.,
2014; Umezawa et al., 2014). Indeed, comparing the results of top-down
inversions to EDGARv4.2 inventory (maximum of bottom-up estimates for China in
Fig. 7), fossil fuel emissions are reduced by 33 % from 30 to 20 Tg CH4
yr-1 (range 9–30) and agriculture and waste emissions are reduced by
27 % from 37 to 27 Tg CH4 yr-1 (range 16–37). This result
is consistent with a new inventory for methane emissions from China based on
county-scale data (43 ± 6 Tg yr-1),
indicating that coal-related methane emissions are 37 % (−7 Tg yr-1)
lower than reported in the EDGARv4.2 inventory (Peng et al.,2016) (see also
Sect. 3.1.2). Thompson et al. (2015) showed that their prior (based on
EDGARv4.2) overestimated the Chinese methane emissions by 30 %; however, they
found no significant difference in the coal sector estimates between prior and
posterior and attribute the difference to rice emissions. It demonstrates that
inversions are capable of verifying regional emissions when biases in the
inventories are substantial, as in the case of China.
In contrast to the Chinese estimates, emissions inferred for Africa and
especially southern Africa are significantly larger than in the prior estimates
(Fig. 6). For example, for southern Africa, the mean of the inversion ensemble
is 44 Tg CH4 yr−1 [37–53], starting at a
mean prior of 36 Tg CH4 yr-1 [27–35]. This is a 25 %
increase compared to mean prior estimates for
southern Africa. A t test
performed for the available estimates suggests that the mean posterior for
southern Africa is different from the prior at the 98 % confidence level. An
increase of northern African emissions is also inferred from the ensemble of
inversions but is less significant.
For all other regions, emission changes compared to prior values remain
within the first and third quartiles of the distributions. In particular,
contiguous USA (without Alaska) is found to emit 41 Tg CH4 yr-1
[34–49], which is close to the prior estimates. Top-down and bottom-up
estimates are consistent for anthropogenic sources in this region. Only natural
wetlands are lower as estimated by top-down models (9 Tg CH4 yr-1
[6–13]) than by bottom-up models (13 Tg CH4 yr-1 [6–23]).
|
|
| 6.0.0.0 | Future developments, missing elements and remaining uncertainties |
| 6.0.0.0 | Future developments, missing elements and remaining uncertainties | | | | Kirschke et al. (2013) identified four main shortcomings in the assessment
of regional to global CH4 budgets, which we revisit now.
Annual to decadal CH4 emissions from natural sources (wetlands,
fresh water, geological) are highly uncertain. The work by Poulter et al.
(2016), following Melton et al. (2013) allows partitioning the uncertainty
(expressed as the range in the estimates) of methane emissions from natural
wetlands between wetland extent and other components, based on the use of a
common and newly developed database for wetland extent. This approach confirms
that wetland extent dominates the uncertainty of modelled methane emissions
from wetlands (30–40 % of the uncertainty). The rest of the uncertainty lies in
the model parameterizations of the flux density, which remains poorly
constrained due to very few methane flux measurements available for different
ecosystems over time. More measurements of the isotopic atmospheric composition
of the various ecosystems (bogs/swamps, C3/C4 vegetation, etc.) would also help
better constrain methane fluxes as well as its isotopic signature in the
wetland models. In addition, the footprints of flux measurements are largely on
too small scales (e.g. chamber measurements) to be compared with the lower resolution
at which land surface models operate. Although more and more flux sites now
integrate measurements of methane fluxes by eddy covariance, such a technique
can reveal unexpected issues (e.g. Baldocchi et al., 2012). There is a need for
integration of methane flux measurements on the model of the FLUXNET activity (http://fluxnet.ornl.gov/).
This would allow further refinement of the model parameterizations (Turetsky et
al., 2014; Glagolev et al., 2011). A comparison of the model ensemble estimates
against bottom-up inventory for western Siberia by Glagolev et al. (2011) made
by Bohn et al. (2015) showed that there still is a sizable disagreement between
their results. A more complete analysis of the literature for freshwater
emissions has led to a 50 % increase of the reported range compared to Kirschke
et al. (2013). Emitting pathways such as ebullition remain poorly understood
and quantified. There is a need for systematic measurements from a suite of
sites reflecting the diversity of lake morphologies to better understand the
short-term biological control on ebullition variability (Wik et al., 2014).
Similarly more local measurements using continuous-laser-based techniques would
allow refining the estimation of geological methane emissions. Further efforts
are needed: (1) extending the monitoring of the methane emissions from the
different natural sources (wetlands, fresh waters and geological) complemented
with key environmental variables to allow proper interpretation (e.g. soil
temperature and moisture, vegetation types, water temperature, acidity,
nutrient concentrations, NPP, soil carbon density); (2) developing
process-based modelling approaches to estimate inland emissions instead of
data-driven extrapolations of unevenly distributed and local flux observations;
and (3) creating a global flux product for all inland water emissions at high
resolution allowing the avoidance of double counting between wetlands and
freshwater systems.
The partitioning of CH4 emissions and sinks by region and process
is not sufficiently constrained by atmospheric observations in top-down models.
In this work, we report inversions assimilating satellite data from GOSAT (and
one inversion using SCIAMACHY), which bring more constraints, especially over
tropical continents. The extension of the CH4 surface networks to
poorly observed regions (e.g. tropics, China, India, high latitudes) is still
critical to complement satellite data, which do not observe well in cloudy
regions and at high latitudes but also to evaluate and correct satellite
biases. Such data now exist for China (Fang et al., 2015), India (Tiwari and
Kumar, 2012; Lin et al., 2015) and Siberia (Sasakawa et al., 2010; Winderlich
et al., 2010) and can be assimilated in inversions in the upcoming years.
Observations from other tracers could help partition the different methane
emitting processes. Carbon monoxide (FortemsCheiney et al., 2011) can provide
constraints for biomass burning for instance. However, additional tracers can
also bring contradictory trends in emissions such as the ones suggested since
2007 by 13C (Schaefer et al., 2016) and ethane (Hausmann et al.,
2016). Such discrepancies have to be understood and solved to be able to
properly use additional tracers to constrain methane emissions. An update of OH
fields is expected in 2016 with an ensemble of chemistry transport model and
chemistry-climate model simulations in the framework of CCMI (Chemistry-Climate
Model Initiative) spanning the past 3 decades (http://www.met.reading.ac.uk/
ccmi/). The outcome of this experiment will contribute to an improved
representation of the methane sink (Lamarque et al., 2013). The development of
regional components of the global methane budget is also a way to improve
global totals by developing regional top-down and bottom-up approaches. Such
efforts are underway for South and East Asia (Patra et al., 2013; Lin et al.,
2015) and for the Arctic (Bruhwiler et al., 2015), where seasonality (e.g. Zona
et al., 2016, for tundra) and magnitude (e.g. Berchet et al., 2016, for
continental shelves) of methane emissions remain poorly understood.
The ability to allocate observed atmospheric changes to changes of a given
source is limited. Most inverse groups use EDGARv4.2 inventory as a prior,
being the only annual gridded anthropogenic inventory to date. An updated
version of the EDGARv4.2 inventory has been recently released
(EDGARv4.2FT2012), which is very close at a global scale to the extrapolation
performed in this paper based on statistics from BP (http://www.bp.com/) and
on agriculture emissions from FAO (http://faostat3.fao.org). However, the
significant changes in emissions in China (decrease) and Africa (increase)
found in this synthesis strongly suggest the necessity to further revise the
EDGAR inventory, in particular for coal-related emissions (China). Such an
update is an ongoing effort in the EDGAR group. More extensive comparisons and
exchange between the different inventory teams would also favour a path towards
more consistency.
Uncertainties in the modelling of atmospheric transport and chemistry limit
the optimal assimilation of atmospheric observations and increase the
uncertainties of the inversionderived flux estimates. In this work, we gathered
more inversion models than in Kirschke et al. (2013), leading to small to
significant regional differences in the methane budget for 2000–2009. For the
next release, it is important to stabilize the core group of participating
inversions in order not to create artificial changes in the reporting of
uncertainties. More, the recent results of Locatelli et al. (2015), who studied
the sensitivity of inversion results to the representation of atmospheric transport,
suggest that regional changes in the balance of methane emissions between
inversions may be due to different characteristics of the transport models used
here as compared to Kirschke et al. (2013). Indeed, the TRANSCOM experiment
synthesized in Patra et al. (2011) showed a large sensitivity of the
representation of atmospheric transport on methane concentrations in the
atmosphere. As an illustration, in their study, the modelled CH4
budget appeared to depend strongly on the troposphere– stratosphere exchange
rate and thus on the model vertical grid structure and circulation in the lower
stratosphere. These results put pressure to continue to improve atmospheric
transport models, especially on the vertical.
|
|
| 7.0.0.0 | Conclusions |
| 7.0.0.0 | Conclusions | | | | We have built a global methane budget by gathering and synthesizing a large
ensemble of published results using a consistent methodology, including
atmospheric observations and inversions (top-down inversions), process-based
models for land surface emissions and atmospheric chemistry, and inventories of
anthropogenic emissions (bottom-up models and inventories). For the 2003–2012
decade, global methane emissions are 558 Tg CH4 yr-1
(range of 540–568), as estimated by top-down inversions. About 60 % of global
emissions are anthropogenic (range of 50–70 %). Bottom-up models and
inventories suggest much larger global emissions (736 Tg CH4 yr-1
[596–884]) mostly because of larger and more uncertain natural emissions from
inland water systems, natural wetlands and geological leaks. Considering the
atmospheric constraints on the top-down budget, it is likely that some of the
individual emissions reported by the bottomup approaches are overestimated,
leading to too large global emissions from a bottom-up perspective.
The latitudinal breakdown inferred from top-down approaches reveals a
domination of tropical emissions (∼ 64 %) as compared to mid (∼
32 %) and high (∼ 4 %) northern latitudes (above 60◦ N). The
three largest emitting regions (South America, Africa, South East Asia) account
for almost 50 % of the global budget. Top-down inversions consistently infer
lower emissions in China (∼ 58 Tg CH4 yr-1
[51–72]) compared with the EDGARv4.2 inventories (> 70 Tg CH4 yr-1)
but more consistent with the USEPA and GAINS inventories and with a recent
regional inventory (∼ 45 Tg yr-1). On the other hand,
bottom-up methane emissions from Africa are lower than inferred from top-down
inversions. These differences between top-down inversions and inventories call
for a revisit of the emission factors and activity numbers used by the latter,
at least for China and Africa.
Our results, including an extended set of inversions, are compared with the
former synthesis of Kirschke et al. (2013), showing good consistency overall
when comparing the same decade (2000–2009) at the global scale. Significant
differences occur at the regional scale when comparing the 2000– 2009 decadal
emissions. This important result indicates that using different transport models
and inversion setups can significantly change the partition of emissions at the
regional scale, making it less robust. It also means that we need to gather a
stable, and as complete as possible, core of transport models in the next
release of the budget in order to integrate this uncertainty within the budget.
Among the different uncertainties raised in Kirschke et al. (2013), the
present work estimated that 30–40 % of the large range associated with modelled
wetland emissions in Kirschke et al. (2013) was due to the estimation of
wetland extent. The magnitudes and uncertainties of all other natural sources
have been revised and updated, which has led to decreased the emission
estimates for oceans, termites, wild animals and wildfires, and to increased
emission estimates and range for freshwater systems. Although the risk of
double counting emissions between natural and anthropogenic gas leaks exists,
total fossil-related reported emissions are found consistent with atmospheric 14C.
This places a clear priority on reducing uncertainties in emissions from inland
water systems by better quantifying the emission factors of each contributor
(streams, rivers, lakes, ponds) and eliminating the (plausible) double counting
with wetland emissions. The development
of process-based models for inland water emissions, constrained by
observations, is a priority to overcome the present uncertainties on inland
water emissions. Also important, although not addressed here, is to revise and
update the magnitude, regional distribution, interannual variability and
decadal trends in the OH radicals in the troposphere and stratosphere. This
should be possible soon by the release of the CCMI ongoing multimodel
intercomparison (http://www.igacproject.org/CCMI). Our work also suggests the
need for more interactions among groups developing the emission inventories in
order to resolve discrepancies on the magnitude of emissions and trends in key
regions such as China or Africa. Particularly, the budget assessment of these
regions should strongly benefit from the ongoing effort to develop a network of
in situ atmospheric measurement stations. Finally, additional tracers (methane
isotopes, ethane, CO) have potential to bring more constraint on the global
methane cycle if their information content relative to methane emission trends
is consistent with each other, which is not fully the case at present (Schaefer
et al., 2016; Hausmann et al., 2016). Building on the improvement of the points
above, our aim is to update this synthesis as a living review paper on a
regular basis (∼ every 2 years). Each update will produce a more
recent decadal CH4 budget, highlight changes in emissions and
trends, and show the availability and inclusion of new data, as well as model
improvements.
On the top of the decadal methane budget presented in this paper, trends and
year-to-year changes in the methane cycle have been highly discussed in the
recent literature, especially because a sustained atmospheric positive growth
rate of more than >+5 ppb yr-1 has been observed since 2007 after
almost a decade of stagnation in the late 1990s and early 2000s (Dlugokencky et
al., 2011, Nisbet et al., 2014). Scenarios of increasing fossil and/or
microbial sources have been proposed to explain this increase (Bousquet et al.,
2011; Bergamaschi et al., 2013; Nisbet et al., 2014). Whereas the decreasing
trend in δ13C in CH4
suggests a significant, if not dominant, contribution from increasing emissions
by microbial CH4 sources (Schaefer et al., 2016; Nisbet et al.,
2014), concurrent ethane and methane column measurements suggest a significant
role (likely at least 39 %) for oil and gas production (Hausmann et al., 2016),
which could be consistent when assuming a concomitant decrease in biomass
burning emissions (heavy source for 13C), as suggested by the
GFED database (Giglio et al., 2013). Yet accounting for the uncertainties in
the isotopic signatures of the sources and their trends may suggest different
portionings of the global methane sources between fossil fuel and biogenic
methane emissions (Schwietzke et al., 2016). A possible positive OH trend has
occurred since the 1970s followed by stagnation to decreasing OH in the 2000s,
possibly contributing significantly to recent observed atmospheric methane
changes (Dalsøren et al., 2016; Rigby et al., 2008; McNorton et al., 2016). The
challenging increase of atmospheric methane during the past decade needs more
efforts to be fully understood. GCP will take its part in analysing and
synthesizing recent changes in the global to regional methane cycle based on
the ensemble of top-down and bottom-up studies gathered for the budget analysis
presented here.
|
|
| 8.0.0.0 | Data availability |
| 8.0.0.0 | Data availability | | | | The data presented here are made available in the belief that their wide
dissemination will lead to greater understanding and new scientific insights on
the methane budget and its changes and help to reduce the uncertainties in the
methane budget. The free availability of the data does not constitute
permission for publication of the data. For research projects, if the data used
are essential to the work, or if the conclusion or results depend on the data,
co-authorship may need to be considered. Full contact details and information
on how to cite the data are given in the accompanying database.
The accompanying database includes one Excel file organized in the following
spreadsheets and two netcdf files defining the regions used to produce the
regional budget.
The file Global_Methane_Budget_2000-2012_v1.1.xlsx includes (1) a summary,
(2) the methane observed mixing ratio and growth rate from the four global
networks (NOAA, AGAGE, CSIRO and UCI), (3) the evolution of global
anthropogenic methane emissions (excluding biomass burning emissions), used to
produce Fig. 2, (4) the global and regional budgets over 2000–2009 based on
bottom-up approaches, (5) the global and regional budgets over 2000– 2009 based
on top-down approaches, (6) the global and regional budgets over 2003–2012
based on bottom-up approaches, (7) the global and regional budgets over 2003–
2012 based on top-down approaches, (8) the global and regional budgets for year
2012 based on bottom-up approaches, (9) the global and regional budgets for
year 2012 based on top-down approaches, and (10) the list of contributors to
contact for further information on specific data.
This database is available from the Carbon Dioxide Information Analysis
Center (Saunois et al., 2016) and the Global Carbon Project
(http://www.globalcarbonproject.org).
The Supplement related to this article is
available at http://www.earth-syst-sci-data.net/8/697/2016/essd-8-697-2016-supplement.pdf
|
|
|